You are on page 1of 13

BBA - Biomembranes 1860 (2018) 1863–1875

Contents lists available at ScienceDirect

BBA - Biomembranes
journal homepage: www.elsevier.com/locate/bbamem

Review

Membranes as modulators of amyloid protein misfolding and target of T


toxicity☆
⁎ ⁎
Anoop Rawat, Ralf Langen , Jobin Varkey
Zilkha Neurogenetic Institute, University of Southern California, Los Angeles, CA 90033, United States

A R T I C LE I N FO A B S T R A C T

Keywords: Abnormal protein aggregation is a hallmark of various human diseases. α-Synuclein, a protein implicated in
Membrane Parkinson's disease, is found in aggregated form within Lewy bodies that are characteristically observed in the
Amyloid brains of PD patients. Similarly, deposits of aggregated human islet amyloid polypeptide (IAPP) are found in the
Helical pancreatic islets in individuals with type 2 diabetes mellitus. Significant number of studies have focused on how
α-Synuclein
monomeric, disaggregated proteins transition into various amyloid structures leading to identification of a vast
IAPP
number of aggregation promoting molecules and processes over the years. Inasmuch as these factors likely
Aggregation
enhance the formation of toxic, misfolded species, they might act as risk factors in disease. Cellular membranes,
and particularly certain lipids, are considered to be among the major players for aggregation of α-synuclein and
IAPP, and membranes might also be the target of toxicity. Past studies have utilized an array of biophysical tools,
both in vitro and in vivo, to expound the membrane-mediated aggregation. Here, we focus on membrane inter-
action of α-synuclein and IAPP, and how various kinds of membranes catalyze or modulate the aggregation of
these proteins and how, in turn, these proteins disrupt membrane integrity, both in vitro and in vivo. The
membrane interaction and subsequent aggregation has been briefly contrasted to aggregation of α-synuclein and
IAPP in solution. This article is part of a Special Issue entitled: Protein Aggregation and Misfolding at the Cell
Membrane Interface edited by Ayyalusamy Ramamoorthy.

1. Introduction into fibrils. The obvious and imperative question is what triggers the
monomeric form to translate into oligomers and fibrils. The factors that
The ability of a protein to function properly is highly correlated trigger or promote aggregation can be classified as either intrinsic or
with its ability to maintain the native structure(s). Protein homeostasis extrinsic [3]. The intrinsic factors include mutations in the protein,
pathways within cells control and maintain the synthesis, structure and truncations and overexpression of proteins. The extrinsic factors that
degradation of proteins [1]. Any deleterious effect on these pathways have been commonly incriminated are variation in pH, lipid mem-
may culminate in improper conformational changes in a protein or lead branes, posttranslational modifications, metal ions, and other small
to the accumulation of misfolded proteins within a cell. Protein ag- molecules including polyamines and pesticides [3]. In this review, we
gregation due to misfolding is the basis of a multitude of diseases [2]. focus on the role of lipid membranes in modulating the aggregation of
Misfolding of a protein can result in a loss of function, and/or a toxic α-synuclein and IAPP. Here we describe the structure, membrane in-
gain of function. The toxicity of misfolded/aggregated proteins could be teraction and interaction with lipid-like molecules of α-synuclein and
due to, but not limited to, erroneous interaction with other macro- islet amyloid polypeptide (IAPP), and their implications in disease. For
molecules, mislocalization or damage to intracellular membranes. Due comparative purpose, we also discuss aggregation of α-synuclein and
to the apparent nature of aggregated proteins to inflict cellular toxicity, IAPP in solution. Finally, we examine modulation of the membrane
numerous attempts have been made to unravel the underlying me- binding of these proteins as a possible therapeutic strategy in Parkin-
chanism(s) of protein aggregation. A generalized pathway of protein son's disease and type II diabetes.
aggregation is displayed in Fig. 1. Native proteins that have typically
monomeric structure form oligomers, and these ‘on-pathway’ oligomers
eventually form fibrils. Some ‘off-pathway’ oligomers are not converted


This article is part of a Special Issue entitled: Protein Aggregation and Misfolding at the Cell Membrane Interface edited by Ayyalusamy Ramamoorthy.

Corresponding authors.
E-mail addresses: langen@usc.edu (R. Langen), jvarkey@usc.edu (J. Varkey).

https://doi.org/10.1016/j.bbamem.2018.04.011
Received 23 February 2018; Received in revised form 19 April 2018; Accepted 19 April 2018
Available online 25 April 2018
0005-2736/ © 2018 Elsevier B.V. All rights reserved.
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

homeostasis and satiety [30,31].


IAPP derives much of its notoriety from its gain-of-toxic function
that arises from its ability to misfold and ultimately form pancreatic
amyloid aggregates in over 90% of patients with type 2 diabetes
(T2DM) [31–35]. Although the overall etiology of T2DM is complex and
not fully understood, multiple lines of evidence suggest that at least
some species formed during the misfolding process are toxic agents
[31–40]. IAPP toxicity has been observed in a number of cell and an-
imal systems [37,39,41–45]. While there is no complete consensus re-
garding the toxicity of larger aggregates, which are sometimes con-
sidered either toxic or protective, smaller misfolded species that form
early during misfolding are typically thought to be toxic
Fig. 1. Aggregation pathway depicting conversion of native protein into oli- [33–39,46–52]. The link between IAPP misfolding and T2DM patho-
gomers and fibrils. The native protein is typically monomeric. Under the in- physiology is further supported by comparison of IAPP sequences from
fluence of various intrinsic and extrinsic factors the native protein is converted different animals. Several mammals (and some non-mammals) express
to ‘on-pathway’ oligomers that eventually form fibrils or ‘off-pathway’ oligo- IAPP homologs, but not all of them are prone to develop T2DM [31].
mers. Interestingly, species which do not spontaneously develop T2DM, for
example mouse and rat, lack amyloidogenicity in their IAPP owing to
2. Physiological and pathological relevance species-specific sequence variations [53,54]. These animals, however,
could be made diabetic by transgenic expression of human IAPP in their
α-Synuclein has been implicated in development of Parkinson's β-cells [42,44,55,56].
disease (PD) and IAPP in type 2 diabetes mellitus. Although a complete How IAPP, seemingly benign in the secretory granules, misfolds,
understanding of the physiological and pathological roles of these aggregates, and acquires a toxic function has been a focus of extensive
proteins is still lacking, many studies have provided insightful in- research [31,40,57,58]. Evidence suggests a potential mechanism for
formation into the biological significance of these proteins. IAPP toxicity where membranes catalyze IAPP misfolding and where
IAPP, in turn, promotes membrane damage [57–59]. A review of con-
2.1. α-Synuclein formational, structural, and functional consequences of this interplay
between IAPP and membranes follows.
α-Synuclein is a presynaptic protein that is highly expressed in the
brain and implicated in PD. PD is the most prevalent age-related 3. Structures in solution
movement disorder that is characterized by the presence of Lewy bodies
in the neurons of PD patients [4,5]. In PD, the dopaminergic neurons Both α-synuclein and IAPP in their monomeric forms are in-
degenerate, leading to lack of dopamine that is required for proper trinsically disordered (Fig. 2). However, both, due to their structural
neuronal functioning [4]. The underlying mechanism of neuronal de- plasticity, can adopt different conformations and structural assemblies.
generation in PD is unclear. Apart from the observation of aggregated Many studies carried out on these proteins reveal several interesting
α-synuclein in Lewy bodies, several genetic studies have implicated α- structural features, which are summarized below. IAPP and α-synuclein
synuclein in PD [5]. Different α-synuclein missense mutations, A30P, can exist in different monomeric, oligomeric and fibrillar states. All of
E46Q, H50Q, G51D, A53E and A53T, are linked to the autosomal these states possess their own distinct structural features that have been
dominant form of PD [6–10]. Further, α-synuclein gene duplication and studied using different biophysical tools. Depending on the conditions
triplication is also linked to early-onset PD [11,12]. Transgenic animals used, some variations in the structure have been reported.
overexpressing α-synuclein display motor deficit that further corrobo-
rate the role of this protein in PD [13–15]. 3.1. Structure of native α-synuclein
Although this protein was discovered almost two decades ago, the
exact function of α-synuclein is still not fully understood. The initial For a long time, α-synuclein was known as an intrinsically dis-
discovery linked the protein to regulation of neural plasticity [16]. ordered, monomeric protein with residual α-helical propensity
Subsequent studies have associated α-synuclein with functions such as [60–62]. In the recent past, however, reports have suggested that it
regulation of synaptic vesicle recycling, dopamine release, exo- and
endocytosis, vesicular trafficking, interaction with SNARE proteins, and
roles in lipid/fatty acid metabolism and transport among many others
[17–25]. Apart from these functions, α-synuclein is reported to interact
with a wide range of intracellular membranes, which is suggested to be
relevant for its normal physiological functioning. A misregulated or
abnormal membrane interaction, on the other hand, is suggested to
result in neuronal malfunctioning and eventual cell death.

2.2. Islet amyloid polypeptide (IAPP)

IAPP, also called amylin, is a peptide hormone consisting of 37


amino acid residues. Initially produced as an 89-residue preproprotein
in β-cells of the pancreas, it undergoes a series of proteolytic and post-
translational modifications leading to its final mature form [26,27]. Fig. 2. Membrane binding and main amyloidogenic regions of α-synuclein and
Mainly produced by β-cells of the islets of Langerhans, IAPP is co- IAPP. (A) The N-terminal ~100 amino acids of α-synuclein are involved in
packaged and co-secreted with insulin in the secretory granules of the membrane binding while the region between 61–95 (NAC region) is the most
β-cells [28,29]. Although not all tissue targets and physiological func- hydrophobic. (B) The α-helical region of IAPP when bound to membranes is
tions of IAPP are completely known, it is thought to mainly affect the between residues 9–22 and the amyloidogenic region is between residues
liver, gut, and brain, where it might play a role in the control of glucose 20–29.

1864
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

could also exist in multimeric forms such as dimers, trimers, tetramers, Rodriguez et al. used micro-electron diffraction to determine 1.4 Å
and even octamers in its native form [63–66]. The native multimeric structure of microcrystals formed by a segment of α-synuclein con-
forms in these studies are structurally distinct from misfolded oligo- taining residues 68–78, called NACore, which has important role in α-
mers, as they are mainly helical rather than β-sheet in nature [64,67]. synuclein aggregation and cytotoxicity [89]. The structural data re-
Still, questions abound regarding whether the monomeric or multimeric vealed protofibrils formed of pairs of face-to-face β-sheets and a
forms represent the physiological state. It may well be possible that α- structure that had similarity to toxic fibrils of full-length α-synuclein
synuclein exists in a dynamic equilibrium between monomeric and [89]. The core region and the participation of particular amino acids in
different multimeric forms. Alteration of this equilibrium could have the cross-β structure of α-synuclein fibrils depends on the fibril poly-
adverse effects on neuronal functioning [68,69]. morph but, in general, it is between the positions ~30–100. Experi-
mental data also suggest the role of N-terminal region and to some
3.2. Structure of α-synuclein oligomers lesser extent, the C-terminal region, in the formation of fibrils, espe-
cially in the interaction between protofilaments [77,78,85,87,90].
It is often suggested that α-synuclein oligomers are more toxic than Most of the experimental data obtained agree with the presence of
monomers or fibrils. α-Synuclein can form a number of different oli- α-synuclein monomeric units stacked in a parallel in-register arrange-
gomers that are often transiently assembled during aggregation into ment within a protofilament. The β-sheets are arranged anti-parallel to
fibrils, in addition to the ‘off-pathway’ oligomers. Structural studies each other within a single monomeric subunit. The number of β-strands
have been difficult due to the transient, unstable and often hetero- in each monomeric unit varies depending on the fibril polymorph. A
geneous nature of oligomeric species. A few studies, however, have recent study by Tuttle et al., using solid-state NMR spectroscopy, EM
used different methods to isolate or trap the oligomers, enabling and X-ray fiber diffraction, presented a new orthogonal Greek-key to-
structural analysis of some oligomeric forms (for review see [70]). pology consisting of the common amyloid features including parallel,
α-Synuclein oligomers have been categorized using criteria such as in-register β-sheets and hydrophobic-core residues [86]. Further
size, shape, extent and nature of β-sheet and proteinase K digestion structural studies would determine which fibril polymorphs have
[71,72]. Oligomers with morphologies including spherical, annular, structure like the in vivo form.
tubular, circular and elliptical have been observed [70,73–75]. In a
recent study by Fusco et al., two distinct species of stabilized α-synu- 3.4. Structure of IAPP monomer and oligomers
clein oligomers were structurally characterized using solid-state nuclear
magnetic resonance (NMR) [67]. One of them, named type A, was non- Early evidence from circular dichroism (CD) suggested that IAPP
toxic and lacked secondary structure in the rigid regions of the protein, monomers are mainly in random coil conformation with some α-helical
while the toxic type B was found to have considerable amount of β- and β-sheet content [91–95]. This structural organization has been
sheet content in its rigid regions. The N-terminal region of the non-toxic confirmed by a number of studies [96,97], some including solution
oligomer was less dynamic and inaccessible compared to the toxic form NMR, which revealed residual α-helical conformations [98,99]. This
[67]. These structural differences enabled the toxic form to insert more structural feature of IAPP also extends to the physiologically relevant
deeply into lipid bilayers of vesicles mimicking the synaptic vesicle acidic conditions in the secretory vesicles from which it is released
phospholipid composition, causing membrane disruption [67]. [100] (Fig. 3). Under these conditions, IAPP monomers exhibit an α-
Overall, reports suggest that α-synuclein oligomers adopt an anti- helical N-terminal region and disordered C-terminal region [100]. As
parallel β-sheet structure [70,76], whereas fibrils form a parallel ar- described later in more detail for membrane-mediated aggregation, the
rangement [77,78]. The oligomeric core seems to include the same presence of this helical structure also appears to promote IAPP ag-
amino acids that form the fibril core, however, with a different β-sheet gregation in solution [98] (Fig. 2).
arrangement in the latter. The number of monomeric units within dif- Despite being an important amyloid species, high-resolution struc-
ferent oligomeric species mostly range from ~10–90. Although the N- tural studies of oligomers have been difficult, owing to the hetero-
and C-terminal regions of α-synuclein are not part of the fibrillar core, geneity and limited stability of such species. A two-dimensional in-
they could have many intermolecular interactions within the oligomer frared spectroscopy (2D-IR) study observed a parallel β-sheet structure
and, hence, considerable structural remodeling of early oligomeric in- in the region within an oligomeric intermediate [101] previously
teractions is expected for fibril elongation. Various oligomeric species shown to be a disordered loop in mature fibrils [102]. This data sug-
reported by different research groups display significant differences in gests that conversion from β-sheet in oligomer to loop in fibrils poses
their morphology, structure, physicochemical properties and cellular free energy barrier that in turn generates a lag phase in IAPP ag-
toxicity that requires further detailed analysis to know which structure gregation. A recent Fourier transform infrared (FTIR) and Raman
could be pathologically more relevant. spectroscopy study showed considerable α-helical content in early IAPP
oligomers [103]. This study also suggested that such oligomers could
3.3. Structure of α-synuclein fibrils have enhanced membrane affinity.

Many oligomers eventually transform into fibrillar structures. The in 3.5. Structure of IAPP fibrils
vitro fibrils formed by α-synuclein are morphologically very similar to
those isolated from postmortem brains of PD patients [79,80]. Like IAPP fibrils, like those of α-synuclein and many other fibrils, can be
oligomers, α-synuclein fibrils also display large polymorphism that has highly polymorphic [104,105], yet most or all of these polymorphs are
hampered the identification of a consensus fibril structure. Structural largely arranged in parallel, in-register structures, as first determined
studies have been carried out on full-length fibrils generated using by EPR and site-directed spin labeling [106] (Fig. 3). Structural models
varied methodologies and by techniques like solid-state NMR, electron of such parallel, in-register fibrils have been generated by solid-state
paramagnetic resonance, hydrogen/deuterium (H/D) exchange mass NMR, EPR, X-ray diffraction of microcrystals from IAPP fragments and
spectrometry, and electron microscopy [77,78,81–86]. Both twisted computational techniques [102,107–109]. Among these models, the
and straight fibrils with varying diameters have been observed EPR and solid-state NMR models are most similar, both having a
[70,86–88]. Early mapping of the α-synuclein fibril core region has horseshoe type arrangement with two β-strands separated by a bend
come from electron paramagnetic resonance (EPR) experiments, which region. A two-strand model is further supported by data from H/D
also revealed that this region takes up a parallel, in-register structure exchange and 2D-IR spectroscopy [110,111]. Despite their similarities,
[77,78]. These findings were later confirmed by a number of techni- the EPR and NMR models also have some differences. First, there is a
ques, including solid-state NMR [81,85,86]. A recent study by slight variation in the exact lengths of the strands, and the relative

1865
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

Fig. 3. Structural models of human IAPP (IAPP) monomers and fibrils. (A) EPR based model of monomeric IAPP bound to phospholipid membrane. The red-colored
thicker ribbon shows the central helical region (residues 9–20) while the red-colored thinner ribbon (residues 21 and 22) indicates the C-terminal end of the helix.
The N- and C-termini flanking the central helical region are unstructured and depicted with dashed lines. Gold-colored spheres indicate phosphates (Apostolidou
et al. [174]) (B) NMR structure of monomeric IAPP bound to SDS micelles showing α-helical conformation of the peptide (PDB 2L86, Nanga et al. [175]). (C) The
NMR structure of monomeric IAPP in acidic solution (pH 5.3) shows α-helical conformation near the N-terminus (residues C7-F15) while the C-terminal region
remains unstructured (PDB 5MGQ, Rodriguez-Camargo et al. [100]). (D) Solid-state NMR based model of IAPP monomers in fibrils with striated ribbon morphology.
The monomeric subunits have β-hairpin conformations and interact with each other through their C-terminal β-stands (Luca et al. [102]). (E, F, and G) EPR based
structural model of IAPP fibrils. The N- and C-terminal β-strands are separated by a long loop region (Bedrood et al. [107]). (E) Individual monomers (shown as blue,
green, and orange ribbons) in fibrils have considerable stagger. The i to i + 3 contacts between monomers are indicated for the blue monomers. (F) Rotated view of
fibril in (E). (G) The structural model of fibril consisting of 101 monomers shown to emphasize a left-handed helical turn of ~90° between the center of the red boxes.
(For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

positions of the strands with respect to each other are different. While 4. Membrane interaction
the solid-state NMR model has both strands in the same plane (Fig. 3D),
the EPR model is staggered by about 15 Å (Fig. 3E–G). This stagger, as Membrane interaction appears to be an essential feature of α-sy-
per distance measurements, indicates that the two strands in the same nuclein in fulfilling its proposed physiological role. An abnormal
monomer are spaced apart from each other. Interestingly, this model is membrane interaction, on the other hand, is associated with patholo-
analogous to a recent fibril model of Alzheimer's amyloid beta peptide gical consequences for both α-synuclein and IAPP. The pathological
obtained from solid-state NMR and cryo EM [112]. The staggered consequences in the form of membrane damage have been observed for
structure could provide stability and expose large hydrophobic surfaces both α-synuclein and IAPP in numerous in vitro assays and confirmed in
at the fibril ends. It has been suggested that such “sticky ends” could vivo. Various studies to understand the underlying mechanism of these
promote capture of incoming monomers and promote fibril extension interactions and to tease out the role of specific membrane components
[107]. and protein features are discussed below.

1866
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

Fig. 4. Structure of membrane-bound α-sy-


nuclein. (A) The membrane-bound structure
of α-synuclein was determined using electron
paramagnetic resonance. α-Synuclein mainly
adopts an extended-helix on small uni-
lamellar vesicles [115]. (B) NMR structure of
SDS micelle-bound α-synuclein (PDB ID:
1XQ8 [136]). (C) Structure of SLAS micelle-
bound α-synuclein (PDB ID: 2KKW [135])
obtained using NMR and pulsed EPR mea-
surements. α-Synuclein adopts a broken-he-
lical conformation on highly curved SDS and
SLAS micelles.

4.1. Membrane interaction of α-synuclein on phospholipid monolayers surrounding triglyceride-rich lipid dro-
plets in cells treated with fatty acids [65]. However, the disease mu-
α-Synuclein contains seven imperfect 11-amino acid repeats that tants of PD were not as efficient in surrounding the lipid droplets and
have some similarities to those present in apolipoproteins (Fig. 2). The protecting triglyceride turnover [65]. It has been further demonstrated
N-terminal ~100 amino acids that encompass these repeats are the that α-synuclein forms soluble oligomers in cultured mesencephalic
primary mediators for membrane binding. α-Synuclein preferentially cells upon exposing those cells to polyunsaturated fatty acids (PUFAs)
binds to phospholipids with negatively charged headgroups like phos- [128,129]. Since the brains of patients with PD and dementia with
phatidic acid (PA), phosphatidylserine (PS), phosphatidylinositol (PI) Lewy bodies have increased amounts of soluble, lipid-associated αS
and phosphatidylglycerol (PG) [113,114]. It has negligible affinity to- oligomers, the binding of α-synuclein to PUFAs could be important for
wards zwitterionic phospholipids like phosphatidylcholine (PC) and pathology [128,129]. In fact, neuronal cells overexpressing α-synu-
phosphatidylethanolamine (PE) headgroups. The affinity towards an- clein, when exposed to physiological levels of PUFA, result in the for-
ionic phospholipids is due to their interaction with multiple lysine re- mation of proteinaceous Lewy-like inclusions that are preceded by
sidues in the N-terminal membrane binding region of α-synuclein PUFA-induced soluble oligomers. [130]. Further, the N-terminal region
[115]. The binding preference can further extend for a particular ne- of α-synuclein was found, in a study, to be essential for binding to PUFA
gatively charged lipid headgroup over another negatively charged and formation of oligomers [131]. In the same report, a C-terminal
headgroup with same acyl chains [116], or for a particular acyl chain truncation resulted in the acceleration of PUFA-induced α-synuclein
over another acyl chain with same phospholipid headgroup [117]. oligomerization [131]. Further investigation is required to probe any
Thus, the membrane affinity of α-synuclein is modulated by both the link between PUFA-induced soluble oligomers and α-synuclein fi-
headgroup and the acyl chain of a phospholipid molecule. brillization.
Furthermore, α-synuclein binding to membranes is affected by
membrane curvature [113,118]. Studies of the binding affinity to lipid
4.2. Structure of membrane-bound α-synuclein
vesicles of different sizes revealed that the protein binds preferentially
to highly curved membranes [113,118]. The more effective binding to
α-Synuclein readily transforms into a predominantly α-helical
smaller vesicles was attributed to the more prevalent membrane
conformation in the presence of phospholipid membranes
packing defects present in those vesicles. This property correlates well
[113,115,132–134]. The conversion into a helical form also occurs in
with the ability of α-synuclein to bind the relatively small synaptic
the presence of other molecules such as fatty acids or SDS micelles
vesicles (~30 nm diameter).
[117,135,136]. Overall, the N-terminal ~100 amino acids attain a he-
In addition to the influence of lipid properties on α-synuclein
lical conformation while the remaining ~40 amino acids in the C-
membrane binding, some of the familial PD mutations also affect the
terminal region remain unordered [115,133,134,136–138] (Figs. 2 and
membrane binding of α-synuclein [119–123]. It should be noted that all
4). However, different groups have reported structures that largely
the familial mutants discovered so far lie between residues 30 and 53,
differ in whether α-synuclein forms a single extended-helix or a broken-
which is right in the middle of the membrane binding region of α-sy-
helix around a central bend. The location of this bend varies from
nuclein and within the core region of α-synuclein fibrils. The A30P and
structure to structure but is typically around position 40. The solution
A53E mutations seem to reduce the membrane interaction while E46K
NMR structure of α-synuclein bound to SDS micelle showed a bent-
seems to enhance the membrane binding [119,121,123–125]. The
helical structure with two antiparallel helices between 3–37 and 45–92
A53T mutation, on the other hand, does not seem to influence mem-
[136] (Fig. 4B). EPR experiments also revealed the presence of a
brane interaction in any significant way [119].
broken-helical structure on lysophospholipid and sodium lauroyl sar-
Aside from lipids, α-synuclein binds to fatty acids and is proposed to
cosinate (SLAS) micelles [133,135] (Fig. 4C). However, in compre-
transport fatty acids inside cells [126]. α-Synuclein shares sequence
hensive studies using continuous wave EPR and DEER based distance
homology with fatty acid binding proteins. The binding affinity, how-
measurements, in combination with computational refinement, it was
ever, does not seem to be as high as reported for classical fatty acid
shown that α-synuclein forms an extended amphipathic helix on
binding proteins [127]. It was observed that α-synuclein accumulated
phospholipid vesicles [115,132] (Fig. 4A). The formation of an

1867
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

extended-helix was further validated in another EPR study that probed affect the conformation of protein on these lipoprotein nanoparticles or
the structures of α-synuclein on bicelles, vesicles and rod-like micelles the formation of these nanoparticles could have pathological implica-
[134]. The observation of a broken-helical structure on detergent mi- tions. Other structures like that observed on cylindrical micelles
celles using solution NMR could be explained by geometrical con- (membrane tubes) may represent binding of α-synuclein to lipid as-
sideration. A detergent micelle (~5 nm diameter) cannot accommodate semblies, often referred to as the “hemifusion” or “hemifission” states.
the fully-extended helical α-synuclein (~14 nm long) on its surface, and These states are thought to form during membrane fusion and fission
thus, a kink is introduced in the middle to accommodate α-synuclein, during processes such as endocytosis or exocytosis. Since α-synuclein is
resulting in a broken-helical form (Fig. 4). linked to synaptic vesicle recycling, the α-synuclein conformation on
Solid-state NMR has recently been used to study the structural cylindrical micelles might represent a functionally active form during
transition from a predominantly α-helical, membrane-bound con- synaptic vesicle recycling. Regarding pathological implications, any
formation to β-sheet fibrils [139]. The fibrils formed in the absence and imbalance between the monomeric and multimeric forms could trigger
presence of lipids displayed similar morphology. Further analyses, non-physiological binding or protein aggregation. α-Synuclein in its
however, revealed major structural differences in their N-termini and a helical conformation induces membrane remodeling, and excessive
minor structural variation in the central NAC domain [139]. During the membrane curvature induction can lead to membrane disruption. Some
fibril formation in presence of lipids, four major states of α-synuclein lipids, particularly negatively charged lipids or lipids with different
were observed. The first state had an α-helical secondary structure that intrinsic curvature could shift the equilibrium towards more α-synu-
converts to a mostly β-sheet secondary structure with a strong asso- clein binding to the membrane surface, and thereby cause membrane
ciation with lipid acyl chains. This was followed by a third state with disruption. Indeed, this has been observed in case of interaction with
key chemical shift signatures that correlates with the mature fibrils and many cellular membranes [152–156].
a fourth state with further changes in chemical shifts indicated qua-
ternary assembly of protofilaments [139]. 4.3. Membrane interaction of IAPP
Another study focused on understanding the role of different regions
of α-synuclein towards the membrane interaction by exploring the Several studies have indicated that membrane interaction of IAPP
membrane bound structure using solid-state NMR [140]. It was found can have deleterious effects on membranes [37,157–160]. IAPP is
that α-synuclein can be divided into three distinct regions based on normally secreted but it has also been observed cytosolically, where it
structural and dynamical properties [140]. First, an N-terminal helical could promote toxicity [55,161–163]. IAPP has been shown to be toxic
segment that acts as a membrane anchor, followed by a central region to islet β-cells and insulinoma cell lines, regardless of whether it is
that determines the affinity for membranes and acts as a sensor of the exogenously applied or endogenously overexpressed. Moreover, there is
lipid properties, and finally, an unstructured C-terminal region [140]. evidence that exogenously applied IAPP can be taken up into the cell in
The ability of α-synuclein to adopt different binding modes and the a process that can be further modulated by cholesterol [164]. Once in
observation of either a broken or an extended-helix conformation is the cell, IAPP can target and disrupt mitochondrial, lysosomal and
attributed to a small difference in free energy between the two con- other intracellular membranes, and cause toxicity [161,165–171]. In-
formations [115,138,141–143]. terestingly, synthetic molecules impairing membrane binding of IAPP
Considering the relatively small energy differences between various can protect cells against toxicity, further supporting the notion that
membrane or lipid-bound forms of α-synuclein, one might expect that membranes are a target of IAPP cytotoxicity [172].
the structural plasticity allows α-synuclein to take up different con-
formations on membranes that vary in shape and curvature, for ex- 4.4. Structure of membrane-bound IAPP
ample tubules and nanoparticles [115,135,144]. This is extremely im-
portant to keep in mind when carrying out structural studies, as α- The potential role of IAPP membrane interactions in disease
synuclein can remodel membranes into nanoparticles and membrane [37,48,58,59,158,165,173] has fostered interest in this process on a
tubes (micellar tubes and bilayer tubes) by induction of membrane mechanistic and structural level. IAPP aggregation can be significantly
curvature [117,144–150]. Tubulation and formation of lipoprotein enhanced by the presence of membranes. The first evidence for a
nanoparticles are a common property of synucleins and apolipoproteins membrane-bound intermediate that is generated prior to β-sheet fibril
[144,145,151]. The structures of the nanoparticle and tubule-bound formation came from CD studies, which revealed transient formation of
forms of synuclein are distinct. While a broken-helical structure is α-helical structure in the presence of anionic lipids [166]. The transient
present on nanoparticles, an extended-helical structure is formed on nature of this helical structure has complicated some structural studies,
tubes [117,144]. Similar data were obtained for α-synuclein's structure but it was ultimately possible to stabilize it long enough for analyses
on tubes and nanoparticles formed in the presence of oleic acid [144]. using site-directed spin labeling and EPR spectroscopy [174]. This
Overall, the available structural data from the different membrane or study found that the IAPP forms a single α-helix encompassing residues
detergent-bound forms of α-synuclein reveal the common theme that 9–22, with flanking N- and C-terminal regions being largely disordered
the conformation is strongly correlated with the size of the binding (Figs. 2 and 3A) (see more detailed discussion below).
template. α-Synuclein adopts a broken-helical structure on surfaces that NMR studies have used membrane-mimetic detergent micelles to
have smaller dimensions than the length of the extended α-synuclein study IAPP structure [175,176]. These studies focused on human IAPP
helix (~14 nm) (Fig. 4). Such surfaces include detergent micelles in detergent micelles [175,176] where it takes on a more extensive two-
(~5 nm diameter) or nanoparticles (~10 nm diameter). On the other helical structure in the region between residues 5–28 [176] and re-
hand, an extended-helix is strongly favored on surfaces that can ac- sidues 7–28 [175] (Fig. 3B). While the structural models of membrane-
commodate the helix in its extended form, such as tubes or intact bound IAPP [174,177] have substantial similarities, they nonetheless
membranes or vesicles. differ significantly from those obtained for micelles. This is especially
Although the complete significance of these studies to α-synuclein the case with respect to the one helical structure on membranes versus
biology remains to be determined, there are several implications. First, the two-helical structure on detergents (Fig. 3A–B). This scenario is
some conformations of α-synuclein may represent functionally active reminiscent of what was observed for α-synuclein (discussed above,
structures. These structures could be formed in response to the changes Fig. 4) and illustrates that structures of proteins can be different on
in lipid/fatty acid metabolism. For example, considering that α-synu- micelles and on membranes.
clein might play some role in lipid transport and metabolism in vivo, its The structure of membrane-bound IAPP oligomers had long been
apolipoprotein-like binding to lipoprotein nanoparticles in a broken- inaccessible because of rapid aggregation in presence of membranes. A
helical conformation could be of functional relevance. Mutations that recent solution NMR study, however, managed to stabilize IAPP

1868
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

that there is a strong dependence on the exact conditions and results


differ due to the use of different lipids and different protein/lipid ratios
in the respective studies. Inhibition of fibrillization has been observed
under conditions favoring stabilization of the helical form of α-synu-
clein. These conditions typically require a low protein/lipid ratio and
membranes to which α-synuclein has a high affinity. Such membranes
contain a significant fraction of negatively charged lipids, like PS, PG
and GM1 [184–187]. In contrast, the use of net neutral lipids, to which
synuclein only has a marginal affinity, does not affect aggregation ki-
netics [188].
At high protein/lipid ratios, an acceleration in the rate of fi-
brillization was observed in the presence of negatively charged phos-
pholipids [184,187], including physiological membranes isolated from
Fig. 5. Schematic of membrane-mediated aggregation. Native IAPP or α-sy-
the brain, synaptosomal membranes and synaptic vesicles mimicking
nuclein molecules bind to the membrane and adopt helical conformation. The
membranes [188–190]. Apart from the acceleration of aggregation of
reduced dimensionality on the membrane surface increases local concentration
of IAPP or α-synuclein, which facilitates intermolecular interaction through α-synuclein in the intracellular space, a recent work demonstrated ac-
their respective amyloidogenic region. This interaction results in the formation celeration of α-synuclein aggregation by exosomes, and particularly by
of oligomers, and finally, fibrils. The red cylinder represents helical region and ganglioside lipids GM1 or GM3 present in exosomes [191]. Galvagnion
the double-headed arrow shows intermolecular interaction in the amyloido- et al. recently found that high protein/lipid ratios result in a primary
genic region of IAPP and α-synuclein. The C-terminal of α-synuclein is not nucleation rate that is at least three orders of magnitude greater than
depicted and the peptide/protein length is not to scale. (For interpretation of the nucleation in bulk solution [187]. The enhanced nucleation rate on
the references to color in this figure legend, the reader is referred to the web the membrane surface was attributed to the higher local concentration
version of this article.) of protein molecules due to the reduced dimensionality [187] (Fig. 5).
This hypothesis is supported by aggregation of α-synuclein at a nano-
oligomers in the presence of lipid nanodiscs [177]. This structure shows molar or even lower bulk concentration in the presence of surfaces like
three antiparallel β-strands formed by residues A8-L12, F15-H18, and hydrophilic glass and cell membrane-mimicking supported lipid bi-
I26-S29 connected via flexible loops in each monomeric unit [177]. This layers [192,193]. Another aspect to favor nucleation on the membrane
structure is remarkably different from the fibril structure underscoring surface is the possibility of membranes directing α-synuclein into
the structural plasticity of this amyloidogenic peptide. conformations that may favor primary nucleation. The conformations
that favor nucleation would also be present in solution albeit rarer than
at the membrane surface. Many groups have suggested that such nu-
5. Membrane-mediated aggregation
cleation-promoting conformations could be further favored by a re-
duction in the membrane interaction of the region around the hydro-
Lipids can modulate the aggregation behavior of α-synuclein and
phobic core (~60–90 position) of α-synuclein, for example, by familial
IAPP (Fig. 5). Both increases and decreases in the rate of aggregation
mutations like A30P and G51D [143,194–196]. In these mutations, the
have been reported and these effects are highly dependent on the lipid
membrane interaction through their N-terminal region remains un-
compositions and protein-to-lipid ratios.
affected as that region still works as a membrane anchor. On the other
hand, the membrane interaction of the central hydrophobic region is
5.1. Effect of lipids on α-synuclein aggregation affected, and this region then becomes available for lateral protein-
protein interaction on the membrane surface. Such interactions in turn
At physiological pH, α-synuclein aggregates in the test tube into would augment membrane-mediated aggregation of α-synuclein.
fibrils that are morphologically indistinguishable from those found in Apart from point mutations, truncation of the C-terminal region
Lewy bodies [73,87]. However, the fibrillization reaction requires high encompassing several negative charges enhances the aggregation pro-
protein concentrations and incubation times of several days pensity of α-synuclein [90,188,197]. Moreover, C-terminal truncation
[73,74,178]. The speculation that membranes might modulate ag- mutants have been observed in Lewy bodies, thereby signifying the
gregation of α-synuclein originated as lipids were found in the amyloid biological relevance of this α-synuclein fragment [90,197–201].
deposits in different human diseases [179,180]. In Parkinson's disease, Another membrane factor that can influence α-synuclein aggrega-
Lewy bodies (LB) contain aggregated α-synuclein along with other tion is membrane curvature. Smaller vesicles were more efficient in
protein components and a significant proportion of lipids that are increasing the rate of fibrillization, presumably due to the aforemen-
thought to be derived from degraded membrane organelles [181,182]. tioned enhanced binding [184]. This could indicate that highly curved
Membranes influence the aggregation of α-synuclein in two dif- membranes or membrane regions could be ‘hot spots’ for aggregation.
ferent ways: one, by affecting the rate of aggregation, whether it is As described previously, α-synuclein can bend membranes and form
inhibition or acceleration, and two, by becoming incorporated into lipoprotein nanoparticles and membrane tubes. The lipoprotein parti-
lipid-protein co-aggregates/fibrils. The co-aggregation occurs by the cles have multiple α-synuclein molecules coming together in close
‘pulling-out’ of the lipids from the membrane during the aggregation of proximity [144]. Similarly, the membrane tubes are formed at high
α-synuclein on the membrane surface. A recent in vitro study found that protein/lipid ratios and α-synuclein molecules are likely to be locally
this is particularly prevalent in case of membranes containing nega- concentrated on tubes. These close contacts on lipoprotein nano-
tively charged lipids [183]. A close association between lipids and α- particles or membrane tubes could aid nucleation of α-synuclein mo-
synuclein was observed at high protein/lipid ratios while at low ratios, lecules and progress them towards the aggregation pathway. Con-
lipid vesicles adsorbed along the fibrils. The process of co-aggregation centration of α-synuclein on highly curved membrane regions within
generated lipid-protein assemblies that had distinct structure, dynamics the cells could, thus, be a mechanism of aggregation. In this case,
and morphology as compared to structures formed by either lipid or membrane remodeling would precede protein aggregation. Lipids or
protein alone [183]. fatty acids with intrinsic curvatures that affect the highly curved re-
Regarding the effect of membranes on the rate of α-synuclein ag- gions in the cellular membranes could assist α-synuclein molecules to
gregation, there was some discussion initially whether membranes en- concentrate in the highly curved regions. Whether an alteration in the
hanced or inhibited α-synuclein aggregation [184–186]. It now appears manner these proteins come together on the lipoprotein nanoparticles

1869
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

or on highly curved regions on cellular membranes leads to protein have been reported to be formed by many amyloidogenic proteins, in-
aggregation remains to be tested. cluding α-synuclein and IAPP [157,158,219,220]. Typically, they in-
volve oligomeric structures that span the bilayer and allow passage of
5.2. Effect of lipids on IAPP aggregation certain solutes through the membrane. It is easy to see how such a
disruption of bilayer integrity could adversely affect local concentra-
A number of studies have shown that membranes can significantly tions of ions and other molecules and, ultimately, lead to toxicity. Al-
promote the aggregation of IAPP into β-sheet rich species though the outcome of disruption of membrane integrity is similar, a
[94,166,202,203]. This aggregation is most pronounced in the presence detergent-like membrane dissolution and incorporation into growing
of negatively charged membranes [166,204–207], for which the posi- fibrils present an entirely different mechanism as they directly attack
tively charged IAPP has a high affinity. Interestingly, cholesterol and the bilayer integrity rather than by providing a proteinaceous pore in an
gangliosides, which are often found enriched in raft-like membranes, otherwise unperturbed membrane.
significantly modulate IAPP aggregation [208–212]. Presence of metal Membrane disruption is not unique to the misfolded β-sheet forms,
ions such as Ca2+ may further modulate interaction of IAPP with but it is also a property of the α-helical forms of α-synuclein and IAPP.
cholesterol containing vesicles by interfering with the pre-fibrillar Peptides/proteins that do not typically aggregate, like rIAPP [202],
phase [211]. The helical structure of membrane-bound IAPP provided truncated IAPP [221], and β-synuclein [145], as well as IAPP and α-
some first insight into potential aggregation mechanisms [174]. This synuclein in helical form (prior to the formation of β-sheet), are capable
helical structure anchors the peptide to the membrane while leaving the of causing membrane leakage. Recent studies have indicated that the
most amyloidogenic region (residues 20–29) exposed for aggregation membrane damage from the α-helical forms of these proteins can arise
(Fig. 3A). Once on the membrane, IAPP molecules are highly con- from membrane remodeling [145,222]. In this mechanism, the helical
centrated and can collide with each other in a two-dimensional manner form acts as a ‘wedge’ on the membrane surface inducing curvature
(Fig. 5). Thus, the highly amyloidogenic regions of IAPP are likely to effects and causing tubulation, vesiculation and formation of lipopro-
come into frequent contact, allowing them to take up β-sheet con- tein particles/lipid-protein complexes. Like α-synuclein, IAPP can also
taining oligomeric form that ultimately proceeds towards fibril forma- remodel vesicles into tubes, smaller vesicles and non-vesicular protein-
tion. Does such transition happen in cellular environment too? This is lipid complexes [222]. The membrane remodeling effect by both α-
difficult to answer since measuring secondary structure of a peptide in synuclein and IAPP was observed only above a certain protein/lipid
cellular environment, within cellular time-frame, without disrupting ratio implying the requirement of a threshold local protein concentra-
normal physiology, has been beyond the reach of current tools and tion on the membrane surface for curvature induction. Thus, the con-
techniques. certed action of multiple wedges is required to provide enough driving
Of critical importance for the enhancement of aggregation are ne- force to trigger membrane remodeling. This also means that membrane
gatively charged lipids [166]. Thus, in principle, exposure to such lipids remodeling should be more pronounced in cases where the local protein
in a cellular environment could be highly toxic. Under normal cir- concentration is very high. In this context, one might speculate that
cumstances, negatively charged lipids are mainly located on the cyto- oligomers could act like ‘delivery vehicles’ to furnish a high local pro-
solic leaflets of membranes. IAPP has been found in the cytosol, and tein concentration on the membrane surface for curvature induction.
might, therefore, have access to such lipids. The vast majority of IAPP, An overexpression of α-synuclein or IAPP could also provide high
however, is secreted to the extracellular surface, where it would only protein concentration for membrane remodeling effects. The patholo-
encounter the external leaflet of the plasma membrane, which may not gical implication of this mechanism was shown by Boassa et al. in an in
promote aggregation as potently. This scenario could be altered by vivo study using transgenic mice, where overexpressed α-synuclein re-
plasticizers and fatty acids, two risk factors implicated in T2DM [203]. modeled membranous organelle system into highly curved structures at
Both of these molecules are amphiphilic, containing an extensive hy- presynaptic termini [223]. The membrane remodeling by both α-sy-
drophobic region as well as a negatively charged group. A recent study nuclein and IAPP was observed in the presence of negatively charged
found that both molecules strongly partition into uncharged mem- lipids. Any change in the cellular membrane composition that increases
branes where they can enhance the charge density. Intriguingly, the negative charge density due to either altered lipid or fatty acid meta-
enhanced charge density also strongly promotes IAPP aggregation just bolism or due to plasticizers could be a risk factor for membrane dis-
as negatively charged lipids do. The implications of this study are that ruption. In summary, the presence of different structural forms
these risk factors may facilitate IAPP aggregation even on cellular (monomers, oligomers, fibrils) that can interact with cellular mem-
membranes that would otherwise be uncharged [203]. branes differing widely in their lipid compositions suggest concerted
effect of multiple mechanisms for membrane disruption in disease pa-
6. Membrane disruption by α-synuclein and IAPP thology.

Many amyloidogenic proteins, including α-synuclein and IAPP, can 7. Strategies for inhibition of membrane-mediated aggregation
disrupt the integrity of cellular membranes [153–156,213,214]. Several
mechanisms of membrane disruption have been proposed and it is Membrane-mediated aggregation can be inhibited by preventing
important to keep in mind that they are not mutually exclusive. This is membrane interaction or by trapping the helical membrane-bound
perhaps best illustrated by the different temporal phases described for form. Inhibition of membrane interaction was used in the case of
IAPP membrane disruption [173,215–217]. The early phase of mem- squalamine, a natural compound with a net positive charge and a high
brane disruption (leakage) precedes the formation of β-sheet structure, affinity for anionic phospholipids [224]. This compound can easily
while later stages occur around the time of aggregation. Thus, different transport into eukaryotic cells and displace proteins bound to the cy-
structures and properties of IAPP appear to be responsible for com- toplasmic face of plasma membranes [224]. Squalamine was found to
promising membrane integrity. displace α-synuclein from lipid membranes, thereby inhibiting α-sy-
Some commonly suggested mechanisms for membrane damage in- nuclein aggregation by competing for binding sites on vesicles com-
clude pore (or channel) formation, detergent-like membrane dissolu- posed of negatively charged DMPS. Further, the reduction in aggrega-
tion, lipid extraction during co-aggregation, and membrane re- tion of α-synuclein was observed in the PD model of C. elegans with
modeling. Aside from the more recently discovered membrane mitigation of PD-associated paralysis [224]. Endosulfine-alpha (ENSA),
remodeling mechanism, many of the above mentioned mechanisms a protein that is down-regulated in the brains of synucleinopathy pa-
have been described in several reviews before, and will, therefore, be tients acts via the trapping mechanism [225]. It possesses the ability to
only briefly mentioned here (for review see [218]). Pores (or channels) specifically interact with the membrane-bound form of α-synuclein.

1870
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

ENSA was found to inhibit membrane-mediated α-synuclein aggrega- Genetics Study, G51D alpha-synuclein mutation causes a novel parkinsonian-
tion, prevent α-synuclein-induced vesicle disruption and neurotoxicity pyramidal syndrome, Ann. Neurol. 73 (2013) 459–471.
[10] S. Appel-Cresswell, C. Vilarino-Guell, M. Encarnacion, H. Sherman, I. Yu, B. Shah,
[225]. A similar mechanism was used for IAPP by Miranker and col- D. Weir, C. Thompson, C. Szu-Tu, J. Trinh, J.O. Aasly, A. Rajput, A.H. Rajput,
leagues, who designed and synthesized an oligopyridylamide IS5 which A. Jon Stoessl, M.J. Farrer, Alpha-synuclein p.H50Q, a novel pathogenic mutation
binds membrane-bound α-helical intermediate of IAPP and inhibits for Parkinson's disease, Mov. Disord. 28 (2013) 811–813.
[11] M. Farrer, J. Kachergus, L. Forno, S. Lincoln, D.S. Wang, M. Hulihan,
membrane-mediated aggregation and IAPP cytotoxicity in INS-1 cells D. Maraganore, K. Gwinn-Hardy, Z. Wszolek, D. Dickson, J.W. Langston,
[172]. In another approach, screening IAPP-derived peptides for their Comparison of kindreds with parkinsonism and alpha-synuclein genomic multi-
inhibition efficacy in the presence of anionic lipids showed that these plications, Ann. Neurol. 55 (2004) 174–179.
[12] A.B. Singleton, M. Farrer, J. Johnson, A. Singleton, S. Hague, J. Kachergus,
peptides can form hetero-complexes with IAPP and inhibit IAPP fi- M. Hulihan, T. Peuralinna, A. Dutra, R. Nussbaum, S. Lincoln, A. Crawley,
brillation [226]. Using yet another strategy, a water soluble derivative M. Hanson, D. Maraganore, C. Adler, M.R. Cookson, M. Muenter, M. Baptista,
of curcumin, CurDAc, was used to inhibit IAPP aggregation in presence D. Miller, J. Blancato, J. Hardy, K. Gwinn-Hardy, Alpha-synuclein locus triplica-
tion causes Parkinson's disease, Science 302 (2003) 841.
of membranes at stoichiometric concentration [227].
[13] E. Masliah, E. Rockenstein, I. Veinbergs, M. Mallory, M. Hashimoto, A. Takeda,
In summary, the multitude of studies described in this review on α- Y. Sagara, A. Sisk, L. Mucke, Dopaminergic loss and inclusion body formation in
synuclein and IAPP have revealed important lipid and protein de- alpha-synuclein mice: implications for neurodegenerative disorders, Science 287
terminants that influence toxic protein aggregation. It appears that (2000) 1265–1269.
[14] M.B. Feany, W.W. Bender, A Drosophila model of Parkinson's disease, Nature 404
membrane-mediated aggregation is a common feature of all amyloids. (2000) 394–398.
In a recent study, for example, using huntingtin exon-1 (Httex1), which [15] P.O. Fernagut, M.F. Chesselet, Alpha-synuclein and transgenic mouse models,
has a central role in Huntington's disease, it was shown that membranes Neurobiol. Dis. 17 (2004) 123–130.
[16] J.M. George, H. Jin, W.S. Woods, D.F. Clayton, Characterization of a novel protein
can potently accelerate aggregation that finally results in significant regulated during the critical period for song learning in the zebra finch, Neuron 15
membrane damage [228]. The membrane-mediated aggregation of (1995) 361–372.
Httex1 was driven by the N17 domain (N terminus containing 17 amino [17] A. Abeliovich, Y. Schmitz, I. Farinas, D. Choi-Lundberg, W.H. Ho, P.E. Castillo,
N. Shinsky, J.M. Verdugo, M. Armanini, A. Ryan, M. Hynes, H. Phillips, D. Sulzer,
acids) that acts as a helical membrane anchor. Like, α-synuclein and A. Rosenthal, Mice lacking alpha-synuclein display functional deficits in the ni-
IAPP, membrane binding leads to high local protein concentration, grostriatal dopamine system, Neuron 25 (2000) 239–252.
thereby promoting intermolecular collision via two-dimensional diffu- [18] D.E. Cabin, K. Shimazu, D. Murphy, N.B. Cole, W. Gottschalk, K.L. McIlwain,
B. Orrison, A. Chen, C.E. Ellis, R. Paylor, B. Lu, R.L. Nussbaum, Synaptic vesicle
sion, which results in formation of aggregates. Membrane damage is a depletion correlates with attenuated synaptic responses to prolonged repetitive
common theme of all amyloid diseases, and therefore, strategies fo- stimulation in mice lacking alpha-synuclein, J. Neurosci. 22 (2002) 8797–8807.
cused on preventing disruption of cellular membranes could be an ef- [19] M.Y. Golovko, G. Barcelo-Coblijn, P.I. Castagnet, S. Austin, C.K. Combs,
E.J. Murphy, The role of alpha-synuclein in brain lipid metabolism: a downstream
fective therapeutic approach in amyloidogenic diseases.
impact on brain inflammatory response, Mol. Cell. Biochem. 326 (2009) 55–66.
[20] S.J. Lee, H. Jeon, K.V. Kandror, Alpha-synuclein is localized in a subpopulation of
Transparency document rat brain synaptic vesicles, Acta Neurobiol. Exp. (Wars) 68 (2008) 509–515.
[21] D.D. Murphy, S.M. Rueter, J.Q. Trojanowski, V.M. Lee, Synucleins are devel-
opmentally expressed, and alpha-synuclein regulates the size of the presynaptic
The Transparency document associated this article can be found, in vesicular pool in primary hippocampal neurons, J. Neurosci. 20 (2000)
online version. 3214–3220.
[22] V.M. Nemani, W. Lu, V. Berge, K. Nakamura, B. Onoa, M.K. Lee, F.A. Chaudhry,
R.A. Nicoll, R.H. Edwards, Increased expression of alpha-synuclein reduces neu-
Acknowledgement rotransmitter release by inhibiting synaptic vesicle reclustering after endocytosis,
Neuron 65 (2010) 66–79.
This work was supported by NIH grant GM115736 and NS084345. [23] T.F. Outeiro, S. Lindquist, Yeast cells provide insight into alpha-synuclein biology
and pathobiology, Science 302 (2003) 1772–1775.
[24] S. Willingham, T.F. Outeiro, M.J. DeVit, S.L. Lindquist, P.J. Muchowski, Yeast
References genes that enhance the toxicity of a mutant huntingtin fragment or alpha-synu-
clein, Science 302 (2003) 1769–1772.
[25] G.S. Withers, J.M. George, G.A. Banker, D.F. Clayton, Delayed localization of sy-
[1] C.L. Klaips, G.G. Jayaraj, F.U. Hartl, Pathways of cellular proteostasis in aging and
nelfin (synuclein, NACP) to presynaptic terminals in cultured rat hippocampal
disease, J. Cell Biol. 217 (2017) 51–63.
neurons, Brain Res. Dev. Brain Res. 99 (1997) 87–94.
[2] F. Chiti, C.M. Dobson, Protein misfolding, amyloid formation, and human disease:
[26] T. Sanke, G.I. Bell, C. Sample, A.H. Rubenstein, D.F. Steiner, An islet amyloid
a summary of progress over the last decade, Annu. Rev. Biochem. 86 (2017)
peptide is derived from an 89-amino acid precursor by proteolytic processing, J.
27–68.
Biol. Chem. 263 (1988) 17243–17246.
[3] L. Breydo, J.M. Redington, V.N. Uversky, Effects of intrinsic and extrinsic factors
[27] L. Marzban, G. Trigo-Gonzalez, C.B. Verchere, Processing of pro-islet amyloid
on aggregation of physiologically important intrinsically disordered proteins, Int.
polypeptide in the constitutive and regulated secretory pathways of beta cells,
Rev. Cell Mol. Biol. 329 (2017) 145–185.
Mol. Endocrinol. 19 (2005) 2154–2163.
[4] J.A. Obeso, M. Stamelou, C.G. Goetz, W. Poewe, A.E. Lang, D. Weintraub, D. Burn,
[28] S.E. Kahn, D.A. D'Alessio, M.W. Schwartz, W.Y. Fujimoto, J.W. Ensinck,
G.M. Halliday, E. Bezard, S. Przedborski, S. Lehericy, D.J. Brooks, J.C. Rothwell,
G.J. Taborsky, D. Porte, Evidence of cosecretion of islet amyloid polypeptide and
M. Hallett, M.R. DeLong, C. Marras, C.M. Tanner, G.W. Ross, J.W. Langston,
insulin by beta-cells, Diabetes 39 (1990) 634–638.
C. Klein, V. Bonifati, J. Jankovic, A.M. Lozano, G. Deuschl, H. Bergman, E. Tolosa,
[29] A. Lukinius, E. Wilander, G.T. Westermark, U. Engström, P. Westermark, Co-lo-
M. Rodriguez-Violante, S. Fahn, R.B. Postuma, D. Berg, K. Marek, D.G. Standaert,
calization of islet amyloid polypeptide and insulin in the B cell secretory granules
D.J. Surmeier, C.W. Olanow, J.H. Kordower, P. Calabresi, A.H.V. Schapira,
of the human pancreatic islets, Diabetologia 32 (1989) 240–244.
A.J. Stoessl, Past, present, and future of Parkinson's disease: a special essay on the
[30] T.A. Lutz, The role of amylin in the control of energy homeostasis, Am. J. Physiol.
200th Anniversary of the Shaking Palsy, Mov. Disord. 32 (2017) 1264–1310.
Regul. Integr. Comp. Physiol. 298 (2010) R1475–1484.
[5] D.J. Moore, A.B. West, V.L. Dawson, T.M. Dawson, Molecular pathophysiology of
[31] P. Westermark, A. Andersson, G.T. Westermark, Islet amyloid polypeptide, islet
Parkinson's disease, Annu. Rev. Neurosci. 28 (2005) 57–87.
amyloid, and diabetes mellitus, Physiol. Rev. 91 (2011) 795–826.
[6] J.J. Zarranz, J. Alegre, J.C. Gomez-Esteban, E. Lezcano, R. Ros, I. Ampuero,
[32] A. Clark, G.J. Cooper, C.E. Lewis, J.F. Morris, A.C. Willis, K.B. Reid, R.C. Turner,
L. Vidal, J. Hoenicka, O. Rodriguez, B. Atares, V. Llorens, E. Gomez Tortosa, T. del
Islet amyloid formed from diabetes-associated peptide may be pathogenic in type-
Ser, D.G. Munoz, J.G. de Yebenes, The new mutation, E46K, of alpha-synuclein
2 diabetes, Lancet 2 (1987) 231–234.
causes Parkinson and Lewy body dementia, Ann. Neurol. 55 (2004) 164–173.
[33] A. Clark, M.R. Nilsson, Islet amyloid: a complication of islet dysfunction or an
[7] M.H. Polymeropoulos, C. Lavedan, E. Leroy, S.E. Ide, A. Dehejia, A. Dutra, B. Pike,
aetiological factor in Type 2 diabetes? Diabetologia 47 (2004) 157–169.
H. Root, J. Rubenstein, R. Boyer, E.S. Stenroos, S. Chandrasekharappa,
[34] R.L. Hull, G.T. Westermark, P. Westermark, S.E. Kahn, Islet amyloid: a critical
A. Athanassiadou, T. Papapetropoulos, W.G. Johnson, A.M. Lazzarini,
entity in the pathogenesis of type 2 diabetes, J. Clin. Endocrinol. Metab. 89 (2004)
R.C. Duvoisin, G. Di Iorio, L.I. Golbe, R.L. Nussbaum, Mutation in the alpha-sy-
3629–3643.
nuclein gene identified in families with Parkinson's disease, Science 276 (1997)
[35] S.E. Kahn, S. Andrikopoulos, C.B. Verchere, Islet amyloid: a long-recognized but
2045–2047.
underappreciated pathological feature of type 2 diabetes, Diabetes 48 (1999)
[8] R. Kruger, W. Kuhn, T. Muller, D. Woitalla, M. Graeber, S. Kosel, H. Przuntek,
241–253.
J.T. Epplen, L. Schols, O. Riess, Ala30Pro mutation in the gene encoding alpha-
[36] L. Haataja, T. Gurlo, C.J. Huang, P.C. Butler, Islet amyloid in type 2 diabetes, and
synuclein in Parkinson's disease, Nat. Genet. 18 (1998) 106–108.
the toxic oligomer hypothesis, Endocr. Rev. 29 (2008) 303–316.
[9] S. Lesage, M. Anheim, F. Letournel, L. Bousset, A. Honore, N. Rozas, L. Pieri,
[37] J. Janson, R.H. Ashley, D. Harrison, S. McIntyre, P.C. Butler, The mechanism of
K. Madiona, A. Durr, R. Melki, C. Verny, A. Brice, G. French Parkinson's Disease
islet amyloid polypeptide toxicity is membrane disruption by intermediate-sized

1871
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

toxic amyloid particles, Diabetes 48 (1999) 491–498. Simone, Structural basis of membrane disruption and cellular toxicity by alpha-
[38] C.A. Jurgens, M.N. Toukatly, C.L. Fligner, J. Udayasankar, S.L. Subramanian, synuclein oligomers, Science 358 (2017) 1440–1443.
S. Zraika, K. Aston-Mourney, D.B. Carr, P. Westermark, G.T. Westermark, [68] U. Dettmer, A.J. Newman, F. Soldner, E.S. Luth, N.C. Kim, V.E. von Saucken,
S.E. Kahn, R.L. Hull, β-Cell loss and β-cell apoptosis in human type 2 diabetes are J.B. Sanderson, R. Jaenisch, T. Bartels, D. Selkoe, Parkinson-causing alpha-synu-
related to islet amyloid deposition, Am. J. Pathol. 178 (2011) 2632–2640. clein missense mutations shift native tetramers to monomers as a mechanism for
[39] A. Lorenzo, B. Razzaboni, G.C. Weir, B.A. Yankner, Pancreatic islet cell toxicity of disease initiation, Nat. Commun. 6 (2015) 7314.
amylin associated with type-2 diabetes mellitus, Nature 368 (1994) 756–760. [69] U. Dettmer, N. Ramalingam, V.E. von Saucken, T.E. Kim, A.J. Newman, E. Terry-
[40] A. Mukherjee, D. Morales-Scheihing, P.C. Butler, C. Soto, Type 2 diabetes as a Kantor, S. Nuber, M. Ericsson, S. Fanning, T. Bartels, S. Lindquist, O.A. Levy,
protein misfolding disease, Trends Mol. Med. 21 (2015) 439–449. D. Selkoe, Loss of native alpha-synuclein multimerization by strategically mu-
[41] H.J. Hiddinga, N.L. Eberhardt, Intracellular amyloidogenesis by human islet tating its amphipathic helix causes abnormal vesicle interactions in neuronal cells,
amyloid polypeptide induces apoptosis in COS-1 cells, Am. J. Pathol. 154 (1999) Hum. Mol. Genet. 26 (2017) 3466–3481.
1077–1088. [70] N. Cremades, S.W. Chen, C.M. Dobson, Structural characteristics of alpha-synu-
[42] J.W. Höppener, C. Oosterwijk, M.G. Nieuwenhuis, G. Posthuma, J.H. Thijssen, clein oligomers, Int. Rev. Cell Mol. Biol. 329 (2017) 79–143.
T.M. Vroom, B. Ahrén, C.J. Lips, Extensive islet amyloid formation is induced by [71] S.W. Chen, S. Drakulic, E. Deas, M. Ouberai, F.A. Aprile, R. Arranz, S. Ness,
development of Type II diabetes mellitus and contributes to its progression: pa- C. Roodveldt, T. Guilliams, E.J. De-Genst, D. Klenerman, N.W. Wood,
thogenesis of diabetes in a mouse model, Diabetologia 42 (1999) 427–434. T.P. Knowles, C. Alfonso, G. Rivas, A.Y. Abramov, J.M. Valpuesta, C.M. Dobson,
[43] B. Konarkowska, J.F. Aitken, J. Kistler, S. Zhang, G.J.S. Cooper, The aggregation N. Cremades, Structural characterization of toxic oligomers that are kinetically
potential of human amylin determines its cytotoxicity towards islet beta-cells, trapped during alpha-synuclein fibril formation, Proc. Natl. Acad. Sci. U. S. A. 112
FEBS J. 273 (2006) 3614–3624. (2015) E1994–2003.
[44] A.V. Matveyenko, P.C. Butler, Islet amyloid polypeptide (IAPP) transgenic rodents [72] N. Cremades, S.I. Cohen, E. Deas, A.Y. Abramov, A.Y. Chen, A. Orte, M. Sandal,
as models for type 2 diabetes, ILAR J. 47 (2006) 225–233. R.W. Clarke, P. Dunne, F.A. Aprile, C.W. Bertoncini, N.W. Wood, T.P. Knowles,
[45] E.L. Saafi, B. Konarkowska, S. Zhang, J. Kistler, G.J. Cooper, Ultrastructural evi- C.M. Dobson, D. Klenerman, Direct observation of the interconversion of normal
dence that apoptosis is the mechanism by which human amylin evokes death in and toxic forms of alpha-synuclein, Cell 149 (2012) 1048–1059.
RINm5F pancreatic islet beta-cells, Cell Biol. Int. 25 (2001) 339–350. [73] K.A. Conway, J.D. Harper, P.T. Lansbury, Accelerated in vitro fibril formation by a
[46] A. Abedini, A.M. Schmidt, Mechanisms of islet amyloidosis toxicity in type 2 mutant alpha-synuclein linked to early-onset Parkinson disease, Nat. Med. 4
diabetes, FEBS Lett. 587 (2013) 1119–1127. (1998) 1318–1320.
[47] P. Cao, A. Abedini, D.P. Raleigh, Aggregation of islet amyloid polypeptide: from [74] K.A. Conway, S.J. Lee, J.C. Rochet, T.T. Ding, R.E. Williamson, P.T. Lansbury Jr.,
physical chemistry to cell biology, Curr. Opin. Struct. Biol. 23 (2013) 82–89. Acceleration of oligomerization, not fibrillization, is a shared property of both
[48] P. Cao, P. Marek, H. Noor, V. Patsalo, L.-H. Tu, H. Wang, A. Abedini, D.P. Raleigh, alpha-synuclein mutations linked to early-onset Parkinson's disease: implications
Islet amyloid: from fundamental biophysics to mechanisms of cytotoxicity, FEBS for pathogenesis and therapy, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 571–576.
Lett. 587 (2013) 1106–1118. [75] H.A. Lashuel, B.M. Petre, J. Wall, M. Simon, R.J. Nowak, T. Walz, P.T. Lansbury
[49] M.S. Fernández, Human IAPP amyloidogenic properties and pancreatic β-cell Jr., Alpha-synuclein, especially the Parkinson's disease-associated mutants, forms
death, Cell Calcium 56 (2014) 416–427. pore-like annular and tubular protofibrils, J. Mol. Biol. 322 (2002) 1089–1102.
[50] J.W.M. Höppener, C.J.M. Lips, Role of islet amyloid in type 2 diabetes mellitus, [76] M.S. Celej, R. Sarroukh, E. Goormaghtigh, G.D. Fidelio, J.M. Ruysschaert,
Int. J. Biochem. Cell Biol. 38 (2006) 726–736. V. Raussens, Toxic prefibrillar alpha-synuclein amyloid oligomers adopt a dis-
[51] E.T. Jaikaran, A. Clark, Islet amyloid and type 2 diabetes: from molecular mis- tinctive antiparallel beta-sheet structure, Biochem. J. 443 (2012) 719–726.
folding to islet pathophysiology, Biochim. Biophys. Acta 1537 (2001) 179–203. [77] M. Chen, M. Margittai, J. Chen, R. Langen, Investigation of alpha-synuclein fibril
[52] A. Kapurniotu, Amyloidogenicity and cytotoxicity of islet amyloid polypeptide, structure by site-directed spin labeling, J. Biol. Chem. 282 (2007) 24970–24979.
Biopolymers 60 (2001) 438–459. [78] A. Der-Sarkissian, C.C. Jao, J. Chen, R. Langen, Structural organization of alpha-
[53] C. Betsholtz, L. Christmansson, U. Engström, F. Rorsman, V. Svensson, synuclein fibrils studied by site-directed spin labeling, J. Biol. Chem. 278 (2003)
K.H. Johnson, P. Westermark, Sequence divergence in a specific region of islet 37530–37535.
amyloid polypeptide (IAPP) explains differences in islet amyloid formation be- [79] R.A. Crowther, S.E. Daniel, M. Goedert, Characterisation of isolated alpha-synu-
tween species, FEBS Lett. 251 (1989) 261–264. clein filaments from substantia nigra of Parkinson's disease brain, Neurosci. Lett.
[54] P. Westermark, U. Engström, K.H. Johnson, G.T. Westermark, C. Betsholtz, Islet 292 (2000) 128–130.
amyloid polypeptide: pinpointing amino acid residues linked to amyloid fibril [80] L.C. Serpell, J. Berriman, R. Jakes, M. Goedert, R.A. Crowther, Fiber diffraction of
formation, Proc. Natl. Acad. Sci. U. S. A. 87 (1990) 5036–5040. synthetic alpha-synuclein filaments shows amyloid-like cross-beta conformation,
[55] J. Janson, W.C. Soeller, P.C. Roche, R.T. Nelson, A.J. Torchia, D.K. Kreutter, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 4897–4902.
P.C. Butler, Spontaneous diabetes mellitus in transgenic mice expressing human [81] A.D. Dearborn, J.S. Wall, N. Cheng, J.B. Heymann, A.V. Kajava, J. Varkey,
islet amyloid polypeptide, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 7283–7288. R. Langen, A.C. Steven, Alpha-synuclein amyloid fibrils with two entwined,
[56] J.F. Paulsson, A. Andersson, P. Westermark, G.T. Westermark, Intracellular amy- asymmetrically associated protofibrils, J. Biol. Chem. 291 (2016) 2310–2318.
loid-like deposits contain unprocessed pro-islet amyloid polypeptide (proIAPP) in [82] C. Del Mar, E.A. Greenbaum, L. Mayne, S.W. Englander, V.L. Woods Jr., Structure
beta cells of transgenic mice overexpressing the gene for human IAPP and trans- and properties of alpha-synuclein and other amyloids determined at the amino
planted human islets, Diabetologia 49 (2006) 1237–1246. acid level, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 15477–15482.
[57] R. Akter, P. Cao, H. Noor, Z. Ridgway, L.-H. Tu, H. Wang, A.G. Wong, X. Zhang, [83] G. Comellas, L.R. Lemkau, A.J. Nieuwkoop, K.D. Kloepper, D.T. Ladror, R. Ebisu,
A. Abedini, A.M. Schmidt, D.P. Raleigh, Islet amyloid polypeptide: structure, W.S. Woods, A.S. Lipton, J.M. George, C.M. Rienstra, Structured regions of alpha-
function, and pathophysiology, J. Diabetes Res. 2016 (2016) 2798269. synuclein fibrils include the early-onset Parkinson's disease mutation sites, J. Mol.
[58] K. Sasahara, Membrane-mediated amyloid deposition of human islet amyloid Biol. 411 (2011) 881–895.
polypeptide, Biophys. Rev. 10 (2017) 453–462. [84] J. Gath, B. Habenstein, L. Bousset, R. Melki, B.H. Meier, A. Bockmann, Solid-state
[59] S.A. Jayasinghe, R. Langen, Membrane interaction of islet amyloid polypeptide, NMR sequential assignments of alpha-synuclein, Biomol. NMR Assign. 6 (2012)
Biochim. Biophys. Acta 1768 (2007) 2002–2009. 51–55.
[60] B. Fauvet, M.K. Mbefo, M.B. Fares, C. Desobry, S. Michael, M.T. Ardah, E. Tsika, [85] M. Vilar, H.T. Chou, T. Luhrs, S.K. Maji, D. Riek-Loher, R. Verel, G. Manning,
P. Coune, M. Prudent, N. Lion, D. Eliezer, D.J. Moore, B. Schneider, P. Aebischer, H. Stahlberg, R. Riek, The fold of alpha-synuclein fibrils, Proc. Natl. Acad. Sci. U.
O.M. El-Agnaf, E. Masliah, H.A. Lashuel, Alpha-synuclein in central nervous S. A. 105 (2008) 8637–8642.
system and from erythrocytes, mammalian cells, and Escherichia coli exists pre- [86] M.D. Tuttle, G. Comellas, A.J. Nieuwkoop, D.J. Covell, D.A. Berthold,
dominantly as disordered monomer, J. Biol. Chem. 287 (2012) 15345–15364. K.D. Kloepper, J.M. Courtney, J.K. Kim, A.M. Barclay, A. Kendall, W. Wan,
[61] F.X. Theillet, A. Binolfi, B. Bekei, A. Martorana, H.M. Rose, M. Stuiver, S. Verzini, G. Stubbs, C.D. Schwieters, V.M. Lee, J.M. George, C.M. Rienstra, Solid-state NMR
D. Lorenz, M. van Rossum, D. Goldfarb, P. Selenko, Structural disorder of mono- structure of a pathogenic fibril of full-length human alpha-synuclein, Nat. Struct.
meric alpha-synuclein persists in mammalian cells, Nature 530 (2016) 45–50. Mol. Biol. 23 (2016) 409–415.
[62] P.H. Weinreb, W. Zhen, A.W. Poon, K.A. Conway, P.T. Lansbury Jr., NACP, a [87] K.A. Conway, J.D. Harper, P.T. Lansbury Jr., Fibrils formed in vitro from alpha-
protein implicated in Alzheimer's disease and learning, is natively unfolded, synuclein and two mutant forms linked to Parkinson's disease are typical amyloid,
Biochemistry 35 (1996) 13709–13715. Biochemistry 39 (2000) 2552–2563.
[63] W. Wang, I. Perovic, J. Chittuluru, A. Kaganovich, L.T. Nguyen, J. Liao, [88] M.G. Spillantini, R.A. Crowther, R. Jakes, N.J. Cairns, P.L. Lantos, M. Goedert,
J.R. Auclair, D. Johnson, A. Landeru, A.K. Simorellis, S. Ju, M.R. Cookson, Filamentous alpha-synuclein inclusions link multiple system atrophy with
F.J. Asturias, J.N. Agar, B.N. Webb, C. Kang, D. Ringe, G.A. Petsko, Parkinson's disease and dementia with Lewy bodies, Neurosci. Lett. 251 (1998)
T.C. Pochapsky, Q.Q. Hoang, A soluble alpha-synuclein construct forms a dynamic 205–208.
tetramer, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 17797–17802. [89] J.A. Rodriguez, M.I. Ivanova, M.R. Sawaya, D. Cascio, F.E. Reyes, D. Shi,
[64] T. Bartels, J.G. Choi, D.J. Selkoe, Alpha-synuclein occurs physiologically as a S. Sangwan, E.L. Guenther, L.M. Johnson, M. Zhang, L. Jiang, M.A. Arbing,
helically folded tetramer that resists aggregation, Nature 477 (2011) 107–110. B.L. Nannenga, J. Hattne, J. Whitelegge, A.S. Brewster, M. Messerschmidt,
[65] N.B. Cole, D.D. Murphy, T. Grider, S. Rueter, D. Brasaemle, R.L. Nussbaum, Lipid S. Boutet, N.K. Sauter, T. Gonen, D.S. Eisenberg, Structure of the toxic core of
droplet binding and oligomerization properties of the Parkinson's disease protein alpha-synuclein from invisible crystals, Nature 525 (2015) 486–490.
alpha-synuclein, J. Biol. Chem. 277 (2002) 6344–6352. [90] I.V. Murray, B.I. Giasson, S.M. Quinn, V. Koppaka, P.H. Axelsen, H. Ischiropoulos,
[66] J. Burre, M. Sharma, T.C. Sudhof, Alpha-synuclein assembles into higher-order J.Q. Trojanowski, V.M. Lee, Role of alpha-synuclein carboxy-terminus on fibril
multimers upon membrane binding to promote SNARE complex formation, Proc. formation in vitro, Biochemistry 42 (2003) 8530–8540.
Natl. Acad. Sci. U. S. A. 111 (2014) E4274–4283. [91] L.R. McLean, A. Balasubramaniam, Promotion of beta-structure by interaction of
[67] G. Fusco, S.W. Chen, P.T.F. Williamson, R. Cascella, M. Perni, J.A. Jarvis, diabetes associated polypeptide (amylin) with phosphatidylcholine, Biochim.
C. Cecchi, M. Vendruscolo, F. Chiti, N. Cremades, L. Ying, C.M. Dobson, A. De Biophys. Acta 1122 (1992) 317–320.

1872
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

[92] C. Goldsbury, K. Goldie, J. Pellaud, J. Seelig, P. Frey, S.A. Müller, J. Kistler, Biol. 315 (2002) 799–807.
G.J. Cooper, U. Aebi, Amyloid fibril formation from full-length and fragments of [120] E. Jo, J. McLaurin, C.M. Yip, P. St George-Hyslop, P.E. Fraser, Alpha-synuclein
amylin, J. Struct. Biol. 130 (2000) 352–362. membrane interactions and lipid specificity, J. Biol. Chem. 275 (2000)
[93] C.E. Higham, E.T. Jaikaran, P.E. Fraser, M. Gross, A. Clark, Preparation of syn- 34328–34334.
thetic human islet amyloid polypeptide (IAPP) in a stable conformation to enable [121] R.J. Perrin, W.S. Woods, D.F. Clayton, J.M. George, Interaction of human alpha-
study of conversion to amyloid-like fibrils, FEBS Lett. 470 (2000) 55–60. synuclein and Parkinson's disease variants with phospholipids. Structural analysis
[94] S. Jha, D. Sellin, R. Seidel, R. Winter, Amyloidogenic propensities and con- using site-directed mutagenesis, J. Biol. Chem. 275 (2000) 34393–34398.
formational properties of ProIAPP and IAPP in the presence of lipid bilayer [122] P.J. McLean, H. Kawamata, S. Ribich, B.T. Hyman, Membrane association and
membranes, J. Mol. Biol. 389 (2009) 907–920. protein conformation of alpha-synuclein in intact neurons. Effect of Parkinson's
[95] R. Kayed, J. Bernhagen, N. Greenfield, K. Sweimeh, H. Brunner, W. Voelter, disease-linked mutations, J. Biol. Chem. 275 (2000) 8812–8816.
A. Kapurniotu, Conformational transitions of islet amyloid polypeptide (IAPP) in [123] R. Bussell Jr., D. Eliezer, Effects of Parkinson's disease-linked mutations on the
amyloid formation in vitro, J. Mol. Biol. 287 (1999) 781–796. structure of lipid-associated alpha-synuclein, Biochemistry 43 (2004) 4810–4818.
[96] A.S. Reddy, L. Wang, S. Singh, Y.L. Ling, L. Buchanan, M.T. Zanni, J.L. Skinner, [124] P.H. Jensen, M.S. Nielsen, R. Jakes, C.G. Dotti, M. Goedert, Binding of alpha-sy-
J.J. de Pablo, Stable and metastable states of human amylin in solution, Biophys. nuclein to brain vesicles is abolished by familial Parkinson's disease mutation, J.
J. 99 (2010) 2208–2216. Biol. Chem. 273 (1998) 26292–26294.
[97] N.F. Dupuis, C. Wu, J.-E. Shea, M.T. Bowers, Human islet amyloid polypeptide [125] D. Ghosh, S. Sahay, P. Ranjan, S. Salot, G.M. Mohite, P.K. Singh, S. Dwivedi,
monomers form ordered beta-hairpins: a possible direct amyloidogenic precursor, E. Carvalho, R. Banerjee, A. Kumar, S.K. Maji, The newly discovered Parkinson's
J. Am. Chem. Soc. 131 (2009) 18283–18292. disease associated Finnish mutation (A53E) attenuates alpha-synuclein aggrega-
[98] J.A. Williamson, J.P. Loria, A.D. Miranker, Helix stabilization precedes aqueous tion and membrane binding, Biochemistry 53 (2014) 6419–6421.
and bilayer-catalyzed fiber formation in islet amyloid polypeptide, J. Mol. Biol. [126] R. Sharon, M.S. Goldberg, I. Bar-Josef, R.A. Betensky, J. Shen, D.J. Selkoe, Alpha-
393 (2009) 383–396. synuclein occurs in lipid-rich high molecular weight complexes, binds fatty acids,
[99] I.T. Yonemoto, G.J.A. Kroon, H.J. Dyson, W.E. Balch, J.W. Kelly, Amylin pro- and shows homology to the fatty acid-binding proteins, Proc. Natl. Acad. Sci. U. S.
protein processing generates progressively more amyloidogenic peptides that in- A. 98 (2001) 9110–9115.
itially sample the helical state, Biochemistry 47 (2008) 9900–9910. [127] C. Lucke, D.L. Gantz, E. Klimtchuk, J.A. Hamilton, Interactions between fatty acids
[100] D.C. Rodriguez Camargo, K. Tripsianes, K. Buday, A. Franko, C. Göbl, and alpha-synuclein, J. Lipid Res. 47 (2006) 1714–1724.
C. Hartlmüller, R. Sarkar, M. Aichler, G. Mettenleiter, M. Schulz, A. Böddrich, [128] R. Sharon, I. Bar-Joseph, M.P. Frosch, D.M. Walsh, J.A. Hamilton, D.J. Selkoe, The
C. Erck, H. Martens, A.K. Walch, T. Madl, E.E. Wanker, M. Conrad, M.H. de formation of highly soluble oligomers of alpha-synuclein is regulated by fatty acids
Angelis, B. Reif, The redox environment triggers conformational changes and and enhanced in Parkinson's disease, Neuron 37 (2003) 583–595.
aggregation of hIAPP in type II diabetes, Sci. Rep. 7 (2017) 44041. [129] R. Sharon, I. Bar-Joseph, G.E. Mirick, C.N. Serhan, D.J. Selkoe, Altered fatty acid
[101] L.E. Buchanan, E.B. Dunkelberger, H.Q. Tran, P.-N. Cheng, C.-C. Chiu, P. Cao, composition of dopaminergic neurons expressing alpha-synuclein and human
D.P. Raleigh, J.J. de Pablo, J.S. Nowick, M.T. Zanni, Mechanism of IAPP amyloid brains with alpha-synucleinopathies, J. Biol. Chem. 278 (2003) 49874–49881.
fibril formation involves an intermediate with a transient β-sheet, Proc. Natl. [130] K. Assayag, E. Yakunin, V. Loeb, D.J. Selkoe, R. Sharon, Polyunsaturated fatty
Acad. Sci. U. S. A. 110 (2013) 19285–19290. acids induce alpha-synuclein-related pathogenic changes in neuronal cells, Am. J.
[102] S. Luca, W.-M. Yau, R. Leapman, R. Tycko, Peptide conformation and supramo- Pathol. 171 (2007) 2000–2011.
lecular organization in amylin fibrils: constraints from solid-state NMR, [131] H. Karube, M. Sakamoto, S. Arawaka, S. Hara, H. Sato, C.H. Ren, S. Goto,
Biochemistry 46 (2007) 13505–13522. S. Koyama, M. Wada, T. Kawanami, K. Kurita, T. Kato, N-terminal region of alpha-
[103] A. Rawat, B.K. Maity, B. Chandra, S. Maiti, Aggregation-induced conformation synuclein is essential for the fatty acid-induced oligomerization of the molecules,
changes dictate islet amyloid polypeptide (IAPP) membrane affinity, Biochim. FEBS Lett. 582 (2008) 3693–3700.
Biophys. Acta (2018), http://dx.doi.org/10.1016/j.bbamem.2018.03.027. [132] C.C. Jao, A. Der-Sarkissian, J. Chen, R. Langen, Structure of membrane-bound
[104] C. Goldsbury, P. Frey, V. Olivieri, U. Aebi, S.A. Müller, Multiple assembly path- alpha-synuclein studied by site-directed spin labeling, Proc. Natl. Acad. Sci. U. S.
ways underlie amyloid-beta fibril polymorphisms, J. Mol. Biol. 352 (2005) A. 101 (2004) 8331–8336.
282–298. [133] P. Borbat, T.F. Ramlall, J.H. Freed, D. Eliezer, Inter-helix distances in lysopho-
[105] C.S. Goldsbury, G.J. Cooper, K.N. Goldie, S.A. Müller, E.L. Saafi, W.T. Gruijters, spholipid micelle-bound alpha-synuclein from pulsed ESR measurements, J. Am.
M.P. Misur, A. Engel, U. Aebi, J. Kistler, Polymorphic fibrillar assembly of human Chem. Soc. 128 (2006) 10004–10005.
amylin, J. Struct. Biol. 119 (1997) 17–27. [134] E.R. Georgieva, T.F. Ramlall, P.P. Borbat, J.H. Freed, D. Eliezer, Membrane-bound
[106] S.A. Jayasinghe, R. Langen, Identifying structural features of fibrillar islet amyloid alpha-synuclein forms an extended helix: long-distance pulsed ESR measurements
polypeptide using site-directed spin labeling, J. Biol. Chem. 279 (2004) using vesicles, bicelles, and rodlike micelles, J. Am. Chem. Soc. 130 (2008)
48420–48425. 12856–12857.
[107] S. Bedrood, Y. Li, J.M. Isas, B.G. Hegde, U. Baxa, I.S. Haworth, R. Langen, Fibril [135] J.N. Rao, C.C. Jao, B.G. Hegde, R. Langen, T.S. Ulmer, A combinatorial NMR and
structure of human islet amyloid polypeptide, J. Biol. Chem. 287 (2012) EPR approach for evaluating the structural ensemble of partially folded proteins,
5235–5241. J. Am. Chem. Soc. 132 (2010) 8657–8668.
[108] A.V. Kajava, U. Aebi, A.C. Steven, The parallel superpleated beta-structure as a [136] T.S. Ulmer, A. Bax, N.B. Cole, R.L. Nussbaum, Structure and dynamics of micelle-
model for amyloid fibrils of human amylin, J. Mol. Biol. 348 (2005) 247–252. bound human alpha-synuclein, J. Biol. Chem. 280 (2005) 9595–9603.
[109] J.J.W. Wiltzius, S.A. Sievers, M.R. Sawaya, D. Cascio, D. Popov, C. Riekel, [137] C.Y. Cheng, J. Varkey, M.R. Ambroso, R. Langen, S. Han, Hydration dynamics as
D. Eisenberg, Atomic structure of the cross-beta spine of islet amyloid polypeptide an intrinsic ruler for refining protein structure at lipid membrane interfaces, Proc.
(amylin), Protein Sci. 17 (2008) 1467–1474. Natl. Acad. Sci. U. S. A. 110 (2013) 16838–16843.
[110] A.T. Alexandrescu, Amide proton solvent protection in amylin fibrils probed by [138] A.J. Trexler, E. Rhoades, Alpha-synuclein binds large unilamellar vesicles as an
quenched hydrogen exchange NMR, PLoS One 8 (2013) e56467. extended helix, Biochemistry 48 (2009) 2304–2306.
[111] L. Wang, C.T. Middleton, S. Singh, A.S. Reddy, A.M. Woys, D.B. Strasfeld, [139] G. Comellas, L.R. Lemkau, D.H. Zhou, J.M. George, C.M. Rienstra, Structural in-
P. Marek, D.P. Raleigh, J.J. de Pablo, M.T. Zanni, J.L. Skinner, 2DIR spectroscopy termediates during alpha-synuclein fibrillogenesis on phospholipid vesicles, J. Am.
of human amylin fibrils reflects stable β-sheet structure, J. Am. Chem. Soc. 133 Chem. Soc. 134 (2012) 5090–5099.
(2011) 16062–16071. [140] G. Fusco, A. De Simone, T. Gopinath, V. Vostrikov, M. Vendruscolo, C.M. Dobson,
[112] L. Gremer, D. Schölzel, C. Schenk, E. Reinartz, J. Labahn, R.B.G. Ravelli, G. Veglia, Direct observation of the three regions in alpha-synuclein that de-
M. Tusche, C. Lopez-Iglesias, W. Hoyer, H. Heise, D. Willbold, G.F. Schröder, Fibril termine its membrane-bound behaviour, Nat. Commun. 5 (2014) 3827.
structure of amyloid-β(1-42) by cryo-electron microscopy, Science 358 (2017) [141] S.B. Lokappa, T.S. Ulmer, Alpha-synuclein populates both elongated and broken
116–119. helix states on small unilamellar vesicles, J. Biol. Chem. 286 (2011) 21450–21457.
[113] W.S. Davidson, A. Jonas, D.F. Clayton, J.M. George, Stabilization of alpha-synu- [142] A.C. Ferreon, Y. Gambin, E.A. Lemke, A.A. Deniz, Interplay of alpha-synuclein
clein secondary structure upon binding to synthetic membranes, J. Biol. Chem. binding and conformational switching probed by single-molecule fluorescence,
273 (1998) 9443–9449. Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 5645–5650.
[114] M. Ramakrishnan, P.H. Jensen, D. Marsh, Alpha-synuclein association with [143] C.R. Bodner, C.M. Dobson, A. Bax, Multiple tight phospholipid-binding modes of
phosphatidylglycerol probed by lipid spin labels, Biochemistry 42 (2003) alpha-synuclein revealed by solution NMR spectroscopy, J. Mol. Biol. 390 (2009)
12919–12926. 775–790.
[115] C.C. Jao, B.G. Hegde, J. Chen, I.S. Haworth, R. Langen, Structure of membrane- [144] J. Varkey, N. Mizuno, B.G. Hegde, N. Cheng, A.C. Steven, R. Langen, Alpha-sy-
bound alpha-synuclein from site-directed spin labeling and computational re- nuclein oligomers with broken helical conformation form lipoprotein nano-
finement, Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 19666–19671. particles, J. Biol. Chem. 288 (2013) 17620–17630.
[116] G.F. Wang, C. Li, G.J. Pielak, 19F NMR studies of alpha-synuclein-membrane in- [145] J. Varkey, J.M. Isas, N. Mizuno, M.B. Jensen, V.K. Bhatia, C.C. Jao, J. Petrlova,
teractions, Protein Sci. 19 (2010) 1686–1691. J.C. Voss, D.G. Stamou, A.C. Steven, R. Langen, Membrane curvature induction
[117] N. Mizuno, J. Varkey, N.C. Kegulian, B.G. Hegde, N. Cheng, R. Langen, and tubulation are common features of synucleins and apolipoproteins, J. Biol.
A.C. Steven, Remodeling of lipid vesicles into cylindrical micelles by alpha-sy- Chem. 285 (2010) 32486–32493.
nuclein in an extended alpha-helical conformation, J. Biol. Chem. 287 (2012) [146] A.R. Braun, M.M. Lacy, V.C. Ducas, E. Rhoades, J.N. Sachs, Alpha-synuclein-in-
29301–29311. duced membrane remodeling is driven by binding affinity, partition depth, and
[118] B. Nuscher, F. Kamp, T. Mehnert, S. Odoy, C. Haass, P.J. Kahle, K. Beyer, Alpha- interleaflet order asymmetry, J. Am. Chem. Soc. 136 (2014) 9962–9972.
synuclein has a high affinity for packing defects in a bilayer membrane: a ther- [147] C. Eichmann, S. Campioni, J. Kowal, I. Maslennikov, J. Gerez, X. Liu,
modynamics study, J. Biol. Chem. 279 (2004) 21966–21975. J. Verasdonck, N. Nespovitaya, S. Choe, B.H. Meier, P. Picotti, J. Rizo,
[119] E. Jo, N. Fuller, R.P. Rand, P. St George-Hyslop, P.E. Fraser, Defective membrane H. Stahlberg, R. Riek, Preparation and characterization of stable alpha-synuclein
interactions of familial Parkinson's disease mutant A30P alpha-synuclein, J. Mol. lipoprotein particles, J. Biol. Chem. 291 (2016) 8516–8527.

1873
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

[148] M.M. Ouberai, J. Wang, M.J. Swann, C. Galvagnion, T. Guilliams, C.M. Dobson, membrane orientation of IAPP in its natively amidated form at physiological pH in
M.E. Welland, Alpha-synuclein senses lipid packing defects and induces lateral a membrane environment, Biochim. Biophys. Acta 1808 (2011) 2337–2342.
expansion of lipids leading to membrane remodeling, J. Biol. Chem. 288 (2013) [176] S.M. Patil, S. Xu, S.R. Sheftic, A.T. Alexandrescu, Dynamic alpha-helix structure of
20883–20895. micelle-bound human amylin, J. Biol. Chem. 284 (2009) 11982–11991.
[149] A.P. Pandey, F. Haque, J.C. Rochet, J.S. Hovis, Alpha-synuclein-induced tubule [177] D.C. Rodriguez Camargo, K.J. Korshavn, A. Jussupow, K. Raltchev, D. Goricanec,
formation in lipid bilayers, J. Phys. Chem. B 115 (2011) 5886–5893. M. Fleisch, R. Sarkar, K. Xue, M. Aichler, G. Mettenleiter, A.K. Walch, C. Camilloni,
[150] C.H. Westphal, S.S. Chandra, Monomeric synucleins generate membrane curva- F. Hagn, B. Reif, A. Ramamoorthy, Stabilization and structural analysis of a
ture, J. Biol. Chem. 288 (2013) 1829–1840. membrane-associated hIAPP aggregation intermediate, elife 6 (2017).
[151] R.A. Silva, R. Huang, J. Morris, J. Fang, E.O. Gracheva, G. Ren, A. Kontush, [178] M. Necula, C.N. Chirita, J. Kuret, Rapid anionic micelle-mediated alpha-synuclein
W.G. Jerome, K.A. Rye, W.S. Davidson, Structure of apolipoprotein A-I in spherical fibrillization in vitro, J. Biol. Chem. 278 (2003) 46674–46680.
high density lipoproteins of different sizes, Proc. Natl. Acad. Sci. U. S. A. 105 [179] G.P. Gellermann, T.R. Appel, A. Tannert, A. Radestock, P. Hortschansky,
(2008) 12176–12181. V. Schroeckh, C. Leisner, T. Lutkepohl, S. Shtrasburg, C. Rocken, M. Pras,
[152] K. Nakamura, V.M. Nemani, F. Azarbal, G. Skibinski, J.M. Levy, K. Egami, R.P. Linke, S. Diekmann, M. Fandrich, Raft lipids as common components of
L. Munishkina, J. Zhang, B. Gardner, J. Wakabayashi, H. Sesaki, Y. Cheng, human extracellular amyloid fibrils, Proc. Natl. Acad. Sci. U. S. A. 102 (2005)
S. Finkbeiner, R.L. Nussbaum, E. Masliah, R.H. Edwards, Direct membrane asso- 6297–6302.
ciation drives mitochondrial fission by the Parkinson disease-associated protein [180] G.M. Halliday, A. Ophof, M. Broe, P.H. Jensen, E. Kettle, H. Fedorow,
alpha-synuclein, J. Biol. Chem. 286 (2011) 20710–20726. M.I. Cartwright, F.M. Griffiths, C.E. Shepherd, K.L. Double, Alpha-synuclein re-
[153] N. Gosavi, H.J. Lee, J.S. Lee, S. Patel, S.J. Lee, Golgi fragmentation occurs in the distributes to neuromelanin lipid in the substantia nigra early in Parkinson's dis-
cells with prefibrillar alpha-synuclein aggregates and precedes the formation of ease, Brain 128 (2005) 2654–2664.
fibrillar inclusion, J. Biol. Chem. 277 (2002) 48984–48992. [181] K. Wakabayashi, K. Tanji, F. Mori, H. Takahashi, The Lewy body in Parkinson's
[154] G.E. Meredith, S. Totterdell, E. Petroske, K. Santa Cruz, R.C. Callison Jr., Y.S. Lau, disease: molecules implicated in the formation and degradation of alpha-synuclein
Lysosomal malfunction accompanies alpha-synuclein aggregation in a progressive aggregates, Neuropathology 27 (2007) 494–506.
mouse model of Parkinson's disease, Brain Res. 956 (2002) 156–165. [182] K. Wakabayashi, K. Tanji, S. Odagiri, Y. Miki, F. Mori, H. Takahashi, The Lewy
[155] D.D. Song, C.W. Shults, A. Sisk, E. Rockenstein, E. Masliah, Enhanced substantia body in Parkinson's disease and related neurodegenerative disorders, Mol.
nigra mitochondrial pathology in human alpha-synuclein transgenic mice after Neurobiol. 47 (2013) 495–508.
treatment with MPTP, Exp. Neurol. 186 (2004) 158–172. [183] E. Hellstrand, A. Nowacka, D. Topgaard, S. Linse, E. Sparr, Membrane lipid co-
[156] C.C. Stichel, X.R. Zhu, V. Bader, B. Linnartz, S. Schmidt, H. Lubbert, Mono- and aggregation with alpha-synuclein fibrils, PLoS One 8 (2013) e77235.
double-mutant mouse models of Parkinson's disease display severe mitochondrial [184] M. Zhu, A.L. Fink, Lipid binding inhibits alpha-synuclein fibril formation, J. Biol.
damage, Hum. Mol. Genet. 16 (2007) 2377–2393. Chem. 278 (2003) 16873–16877.
[157] M. Anguiano, R.J. Nowak, P.T. Lansbury, Protofibrillar islet amyloid polypeptide [185] V. Narayanan, S. Scarlata, Membrane binding and self-association of alpha-synu-
permeabilizes synthetic vesicles by a pore-like mechanism that may be relevant to cleins, Biochemistry 40 (2001) 9927–9934.
type II diabetes, Biochemistry 41 (2002) 11338–11343. [186] Z. Martinez, M. Zhu, S. Han, A.L. Fink, GM1 specifically interacts with alpha-
[158] T.A. Mirzabekov, M.C. Lin, B.L. Kagan, Pore formation by the cytotoxic islet synuclein and inhibits fibrillation, Biochemistry 46 (2007) 1868–1877.
amyloid peptide amylin, J. Biol. Chem. 271 (1996) 1988–1992. [187] C. Galvagnion, A.K. Buell, G. Meisl, T.C. Michaels, M. Vendruscolo, T.P. Knowles,
[159] R. Kayed, Y. Sokolov, B. Edmonds, T.M. McIntire, S.C. Milton, J.E. Hall, C.M. Dobson, Lipid vesicles trigger alpha-synuclein aggregation by stimulating
C.G. Glabe, Permeabilization of lipid bilayers is a common conformation-depen- primary nucleation, Nat. Chem. Biol. 11 (2015) 229–234.
dent activity of soluble amyloid oligomers in protein misfolding diseases, J. Biol. [188] M.S. Terakawa, Y.H. Lee, M. Kinoshita, Y. Lin, T. Sugiki, N. Fukui, T. Ikenoue,
Chem. 279 (2004) 46363–46366. Y. Kawata, Y. Goto, Membrane-induced initial structure of alpha-synuclein control
[160] P. Westermark, Fine structure of islets of Langerhans in insular amyloidosis, its amyloidogenesis on model membranes, Biochim. Biophys. Acta 1860 (2018)
Virchows Arch. A Pathol. Pathol. Anat. 359 (1973) 1–18. 757–766.
[161] T. Gurlo, S. Ryazantsev, C.-J. Huang, M.W. Yeh, H.A. Reber, O.J. Hines, [189] H.J. Lee, C. Choi, S.J. Lee, Membrane-bound alpha-synuclein has a high ag-
T.D. O'Brien, C.G. Glabe, P.C. Butler, Evidence for proteotoxicity in beta cells in gregation propensity and the ability to seed the aggregation of the cytosolic form,
type 2 diabetes: toxic islet amyloid polypeptide oligomers form intracellularly in J. Biol. Chem. 277 (2002) 671–678.
the secretory pathway, Am. J. Pathol. 176 (2010) 861–869. [190] E. Jo, A.A. Darabie, K. Han, A. Tandon, P.E. Fraser, J. McLaurin, Alpha-synuclein-
[162] P. Westermark, D.L. Eizirik, D.G. Pipeleers, C. Hellerstrom, A. Andersson, Rapid synaptosomal membrane interactions: implications for fibrillogenesis, Eur. J.
deposition of amyloid in human islets transplanted into nude mice, Diabetologia Biochem. 271 (2004) 3180–3189.
38 (1995) 543–549. [191] M. Grey, C.J. Dunning, R. Gaspar, C. Grey, P. Brundin, E. Sparr, S. Linse,
[163] K. Yagui, T. Yamaguchi, A. Kanatsuka, F. Shimada, C.I. Huang, Y. Tokuyama, Acceleration of alpha-synuclein aggregation by exosomes, J. Biol. Chem. 290
H. Ohsawa, K. Yamamura, J. Miyazaki, A. Mikata, et al., Formation of islet amy- (2015) 2969–2982.
loid fibrils in beta-secretory granules of transgenic mice expressing human islet [192] M. Rabe, A. Soragni, N.P. Reynolds, D. Verdes, E. Liverani, R. Riek, S. Seeger, On-
amyloid polypeptide/amylin, Eur. J. Endocrinol. 132 (1995) 487–496. surface aggregation of alpha-synuclein at nanomolar concentrations results in two
[164] S. Trikha, A.M. Jeremic, Clustering and internalization of toxic amylin oligomers distinct growth mechanisms, ACS Chem. Neurosci. 4 (2013) 408–417.
in pancreatic cells require plasma membrane cholesterol, J. Biol. Chem. 286 [193] S. Banerjee, M. Hashemi, Z. Lv, S. Maity, J.C. Rochet, Y.L. Lyubchenko, A novel
(2011) 36086–36097. pathway for amyloids self-assembly in aggregates at nanomolar concentration
[165] J.A. Hebda, A.D. Miranker, The interplay of catalysis and toxicity by amyloid mediated by the interaction with surfaces, Sci. Rep. 7 (2017) 45592.
intermediates on lipid bilayers: insights from type II diabetes, Annu. Rev. Biophys. [194] T. Bartels, L.S. Ahlstrom, A. Leftin, F. Kamp, C. Haass, M.F. Brown, K. Beyer, The
38 (2009) 125–152. N-terminus of the intrinsically disordered protein alpha-synuclein triggers mem-
[166] S.A. Jayasinghe, R. Langen, Lipid membranes modulate the structure of islet brane binding and helix folding, Biophys. J. 99 (2010) 2116–2124.
amyloid polypeptide, Biochemistry 44 (2005) 12113–12119. [195] C.R. Bodner, A.S. Maltsev, C.M. Dobson, A. Bax, Differential phospholipid binding
[167] A. Relini, O. Cavalleri, R. Rolandi, A. Gliozzi, The two-fold aspect of the interplay of alpha-synuclein variants implicated in Parkinson's disease revealed by solution
of amyloidogenic proteins with lipid membranes, Chem. Phys. Lipids 158 NMR spectroscopy, Biochemistry 49 (2010) 862–871.
(2009) 1–9. [196] P. Mazumder, J.E. Suk, T.S. Ulmer, Insight into alpha-synuclein plasticity and
[168] X.-L. Li, T. Chen, Y.-S. Wong, G. Xu, R.-R. Fan, H.-L. Zhao, J.C.N. Chan, misfolding from differential micelle binding, J. Phys. Chem. B 117 (2013)
Involvement of mitochondrial dysfunction in human islet amyloid polypeptide- 11448–11459.
induced apoptosis in INS-1E pancreatic beta cells: an effect attenuated by phy- [197] Y. Izawa, H. Tateno, H. Kameda, K. Hirakawa, K. Hato, H. Yagi, K. Hongo,
cocyanin, Int. J. Biochem. Cell Biol. 43 (2011) 525–534. T. Mizobata, Y. Kawata, Role of C-terminal negative charges and tyrosine residues
[169] M. Magzoub, A.D. Miranker, Concentration-dependent transitions govern the in fibril formation of alpha-synuclein, Brain Behav. 2 (2012) 595–605.
subcellular localization of islet amyloid polypeptide, FASEB J. 26 (2012) [198] W. Hoyer, D. Cherny, V. Subramaniam, T.M. Jovin, Impact of the acidic C-terminal
1228–1238. region comprising amino acids 109–140 on alpha-synuclein aggregation in vitro,
[170] S. Trikha, A.M. Jeremic, Distinct internalization pathways of human amylin Biochemistry 43 (2004) 16233–16242.
monomers and its cytotoxic oligomers in pancreatic cells, PLoS One 8 (2013) [199] K. Levitan, D. Chereau, S.I. Cohen, T.P. Knowles, C.M. Dobson, A.L. Fink,
e73080. J.P. Anderson, J.M. Goldstein, G.L. Millhauser, Conserved C-terminal charge ex-
[171] M. Soty, M. Visa, S. Soriano, M.D.C. Carmona, Á. Nadal, A. Novials, Involvement erts a profound influence on the aggregation rate of alpha-synuclein, J. Mol. Biol.
of ATP-sensitive potassium (K(ATP)) channels in the loss of beta-cell function 411 (2011) 329–333.
induced by human islet amyloid polypeptide, J. Biol. Chem. 286 (2011) [200] W. Li, N. West, E. Colla, O. Pletnikova, J.C. Troncoso, L. Marsh, T.M. Dawson,
40857–40866. P. Jakala, T. Hartmann, D.L. Price, M.K. Lee, Aggregation promoting C-terminal
[172] J.A. Hebda, I. Saraogi, M. Magzoub, A.D. Hamilton, A.D. Miranker, A peptido- truncation of alpha-synuclein is a normal cellular process and is enhanced by the
mimetic approach to targeting pre-amyloidogenic states in type II diabetes, Chem. familial Parkinson's disease-linked mutations, Proc. Natl. Acad. Sci. U. S. A. 102
Biol. 16 (2009) 943–950. (2005) 2162–2167.
[173] P. Cao, A. Abedini, H. Wang, L.-H. Tu, X. Zhang, A.M. Schmidt, D.P. Raleigh, Islet [201] L.A. Volpicelli-Daley, K.C. Luk, T.P. Patel, S.A. Tanik, D.M. Riddle, A. Stieber,
amyloid polypeptide toxicity and membrane interactions, Proc. Natl. Acad. Sci. U. D.F. Meaney, J.Q. Trojanowski, V.M. Lee, Exogenous alpha-synuclein fibrils in-
S. A. 110 (2013) 19279–19284. duce Lewy body pathology leading to synaptic dysfunction and neuron death,
[174] M. Apostolidou, S.A. Jayasinghe, R. Langen, Structure of alpha-helical membrane- Neuron 72 (2011) 57–71.
bound human islet amyloid polypeptide and its implications for membrane- [202] J.D. Knight, J.A. Hebda, A.D. Miranker, Conserved and cooperative assembly of
mediated misfolding, J. Biol. Chem. 283 (2008) 17205–17210. membrane-bound alpha-helical states of islet amyloid polypeptide, Biochemistry
[175] R.P.R. Nanga, J.R. Brender, S. Vivekanandan, A. Ramamoorthy, Structure and 45 (2006) 9496–9508.

1874
A. Rawat et al. BBA - Biomembranes 1860 (2018) 1863–1875

[203] A.K. Okada, K. Teranishi, J.M. Isas, S. Bedrood, R.H. Chow, R. Langen, Diabetic [218] S.M. Butterfield, H.A. Lashuel, Amyloidogenic protein-membrane interactions:
risk factors promote islet amyloid polypeptide misfolding by a common, mem- mechanistic insight from model systems, Angew. Chem. Int. Ed. Engl. 49 (2010)
brane-mediated mechanism, Sci. Rep. 6 (2016) 31094. 5628–5654.
[204] J.D. Knight, A.D. Miranker, Phospholipid catalysis of diabetic amyloid assembly, J. [219] M.J. Volles, P.T. Lansbury Jr., Vesicle permeabilization by protofibrillar alpha-
Mol. Biol. 341 (2004) 1175–1187. synuclein is sensitive to Parkinson's disease-linked mutations and occurs by a pore-
[205] D.H.J. Lopes, A. Meister, A. Gohlke, A. Hauser, A. Blume, R. Winter, Mechanism of like mechanism, Biochemistry 41 (2002) 4595–4602.
islet amyloid polypeptide fibrillation at lipid interfaces studied by infrared re- [220] A. Quist, I. Doudevski, H. Lin, R. Azimova, D. Ng, B. Frangione, B. Kagan, J. Ghiso,
flection absorption spectroscopy, Biophys. J. 93 (2007) 3132–3141. R. Lal, Amyloid ion channels: a common structural link for protein-misfolding
[206] M.F.M. Sciacca, J.R. Brender, D.-K. Lee, A. Ramamoorthy, disease, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 10427–10432.
Phosphatidylethanolamine enhances amyloid fiber-dependent membrane frag- [221] J.R. Brender, E.L. Lee, M.A. Cavitt, A. Gafni, D.G. Steel, A. Ramamoorthy, Amyloid
mentation, Biochemistry 51 (2012) 7676–7684. fiber formation and membrane disruption are separate processes localized in two
[207] X. Zhang, J.R. St Clair, E. London, D.P. Raleigh, Islet amyloid polypeptide mem- distinct regions of IAPP, the type-2-diabetes-related peptide, J. Am. Chem. Soc.
brane interactions: effects of membrane composition, Biochemistry 56 (2017) 130 (2008) 6424–6429.
376–390. [222] N.C. Kegulian, S. Sankhagowit, M. Apostolidou, S.A. Jayasinghe, N. Malmstadt,
[208] W.-J. Cho, S. Trikha, A.M. Jeremic, Cholesterol regulates assembly of human islet P.C. Butler, R. Langen, Membrane curvature-sensing and curvature-inducing ac-
amyloid polypeptide on model membranes, J. Mol. Biol. 393 (2009) 765–775. tivity of islet amyloid polypeptide and its implications for membrane disruption, J.
[209] L. Caillon, L. Duma, O. Lequin, L. Khemtemourian, Cholesterol modulates the Biol. Chem. 290 (2015) 25782–25793.
interaction of the islet amyloid polypeptide with membranes, Mol. Membr. Biol. [223] D. Boassa, M.L. Berlanga, M.A. Yang, M. Terada, J. Hu, E.A. Bushong, M. Hwang,
31 (2014) 239–249. E. Masliah, J.M. George, M.H. Ellisman, Mapping the subcellular distribution of
[210] M.F.M. Sciacca, F. Lolicato, G. Di Mauro, D. Milardi, L. D'Urso, C. Satriano, alpha-synuclein in neurons using genetically encoded probes for correlated light
A. Ramamoorthy, C. La Rosa, The role of cholesterol in driving IAPP-membrane and electron microscopy: implications for Parkinson's disease pathogenesis, J.
interactions, Biophys. J. 111 (2016) 140–151. Neurosci. 33 (2013) 2605–2615.
[211] M.F. Sciacca, D. Milardi, G.M. Messina, G. Marletta, J.R. Brender, [224] M. Perni, C. Galvagnion, A. Maltsev, G. Meisl, M.B. Muller, P.K. Challa,
A. Ramamoorthy, C. La Rosa, Cations as switches of amyloid-mediated membrane J.B. Kirkegaard, P. Flagmeier, S.I. Cohen, R. Cascella, S.W. Chen, R. Limboker,
disruption mechanisms: calcium and IAPP, Biophys. J. 104 (2013) 173–184. P. Sormanni, G.T. Heller, F.A. Aprile, N. Cremades, C. Cecchi, F. Chiti, E.A. Nollen,
[212] M. Wakabayashi, K. Matsuzaki, Ganglioside-induced amyloid formation by human T.P. Knowles, M. Vendruscolo, A. Bax, M. Zasloff, C.M. Dobson, A natural product
islet amyloid polypeptide in lipid rafts, FEBS Lett. 583 (2009) 2854–2858. inhibits the initiation of alpha-synuclein aggregation and suppresses its toxicity,
[213] Y. Porat, S. Kolusheva, R. Jelinek, E. Gazit, The human islet amyloid polypeptide Proc. Natl. Acad. Sci. U. S. A. 114 (2017) E1009–E1017.
forms transient membrane-active prefibrillar assemblies, Biochemistry 42 (2003) [225] D. Ysselstein, B. Dehay, I.M. Costantino, G.P. McCabe, M.P. Frosch, J.M. George,
10971–10977. E. Bezard, J.C. Rochet, Endosulfine-alpha inhibits membrane-induced alpha-sy-
[214] J.R. Brender, U.H.N. Dürr, D. Heyl, M.B. Budarapu, A. Ramamoorthy, Membrane nuclein aggregation and protects against alpha-synuclein neurotoxicity, Acta
fragmentation by an amyloidogenic fragment of human islet amyloid polypeptide Neuropathol. Commun. 5 (2017) 3.
detected by solid-state NMR spectroscopy of membrane nanotubes, Biochim. [226] D. Sellin, L.-M. Yan, A. Kapurniotu, R. Winter, Suppression of IAPP fibrillation at
Biophys. Acta 1768 (2007) 2026–2029. anionic lipid membranes via IAPP-derived amyloid inhibitors and insulin,
[215] S.B. Padrick, A.D. Miranker, Islet amyloid polypeptide: identification of long-range Biophys. Chem. 150 (2010) 73–79.
contacts and local order on the fibrillogenesis pathway, J. Mol. Biol. 308 (2001) [227] A.S. Pithadia, A. Bhunia, R. Sribalan, V. Padmini, C.A. Fierke, A. Ramamoorthy,
783–794. Influence of a curcumin derivative on hIAPP aggregation in the absence and
[216] J.R. Brender, E.L. Lee, K. Hartman, P.T. Wong, A. Ramamoorthy, D.G. Steel, presence of lipid membranes, Chem. Commun. 52 (2016) 942–945.
A. Gafni, Biphasic effects of insulin on islet amyloid polypeptide membrane dis- [228] N.K. Pandey, J.M. Isas, A. Rawat, R.V. Lee, J. Langen, P. Pandey, R. Langen, The
ruption, Biophys. J. 100 (2011) 685–692. 17-residue-long N terminus in huntingtin controls step-wise aggregation in solu-
[217] J.R. Brender, S. Salamekh, A. Ramamoorthy, Membrane disruption and early tion and on membranes via different mechanisms, J. Biol. Chem. 293 (2017)
events in the aggregation of the diabetes related peptide IAPP from a molecular 2597–2605.
perspective, Acc. Chem. Res. 45 (2012) 454–462.

1875

You might also like