Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Angewandte

Research Articles Chemie

How to cite: Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737
Interfaces International Edition: doi.org/10.1002/anie.202011179
German Edition: doi.org/10.1002/ange.202011179

Interaction of a-Synuclein with Phospholipids and the Associated


Restructuring of Interfacial Lipid Water: An Interface-Selective
Vibrational Spectroscopic Study
Biswajit Biswas, Subhadip Roy, Jahur Alam Mondal,* and Prashant Chandra Singh*

Abstract: Interaction of a-Synuclein (aS) with biological (@12) and hydrophilic; whereas the NAC region is highly
lipids is crucial for the onset of its fibrillation at the cell hydrophobic with only one unit negative charge (@1). At the
membrane/water interface. Probed herein is the interaction of physiological concentration ( & 50 mM), aS does not aggre-
aS with membrane-mimicking lipid monolayer/water interfa- gate in bulk aqueous phase (pH & 7.4), even on keeping the
ces. The results depict that aS interacts negligibly with solution for several days.[7] In contrast, aS exhibits accelerated
zwitterionic lipids, but strongly affects the pristine air/water aggregation at membrane-water interfaces (bulk pH
and charged lipid/water interfaces by perturbing the structure & 7.4).[8–10] Aggregation proceeds via pre-nucleation which is
and orientation of the interfacial water. The net negative aS essentially the anchoring of soluble monomeric aS with
(@9 in bulk water; pH 7.4) reorients the water as hydrogen-up membrane lipid head groups at the interface. Pre-nucleation
(H-up) at the air/water interface, and electrostatically interacts controls the subsequent primary and secondary nucleation
with positively charged lipids, making the interface nearly net processes (conformation change) where different sized nuclei
neutral. aS also interacts with negatively charged lipids: the net of aS are formed, leading to the formation of fibrils.[9, 11] The
H-up orientation of the interfacial water decreases at the conformational change of disordered protein at membrane-
anionic lipid/water interface, revealing a domain-specific water interface has been demonstrated using the chiral-VSFG
interaction of net negative aS with the negatively charged technique.[12, 13] Several studies have suggested that the nature
lipids at the membrane surface. of lipid as well as the lipid:aS ratio controls the rate of
aggregation.[7–9, 11] For example, aggregation is more pro-
Introduction nounced at the surface of negatively charged lipids than that

ParkinsonQs and AlzheimerQs diseases are the two impor-


tant neurodegenerative disorders associated with progressive
dementia. The appearance of amyloid plaques in the brain is
the hallmark of such disorders, which in turn linked with the
aggregation of misfolded cellular proteins followed by
fibrillization.[1–3] Hence, understanding the aggregation of
misfolded protein is indispensable for early intervention to
amyloid plaque formation.
a-Synuclein (aS; 140 amino acid residues; Figure 1 a is
one of the disordered proteins whose aggregation is corre-
lated with ParkinsonQs and Lewy body diseases.[4, 5] As shown
in Figure 1 a, the primary structure of aS contains three main
domains, namely N-terminal (blue color), non-amyloid b
component (NAC, ash color), and C-terminal (red color).[6]
The N-terminal contains four units positive charge (+ 4) and
amphipathic in nature; the C-terminal is negatively charged

[*] B. Biswas, Dr. P. C. Singh


School of Chemical Sciences
Indian Association for the Cultivation of Sciences
2A &2B Raja S. C. Mullick Road, Jadavpur, Kolkata 700032 (India)
E-mail: sppcs@iacs.res.in
S. Roy, Dr. J. A. Mondal
Radiation & Photochemistry Division, Bhabha Atomic Research
Centre, Homi Bhabha National Institute Figure 1. a) Amino-acid sequence of aS and the regions of N-terminal,
Trombay, Mumbai 400085 (India) NAC, and C-terminal. Negatively charged residues are shown in red,
E-mail: mondal@barc.gov.in whereas positively charged residues are in blue. b) Chemical structure
Supporting information and the ORCID identification number(s) for of the lipids used: positively charged DPTAP, negatively charged DPPG
the author(s) of this article can be found under: and DPPS, and zwitterionic DPPC. The counter ions with the charged
https://doi.org/10.1002/anie.202011179. lipids are omitted for simplicity.

Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737 T 2020 Wiley-VCH GmbH 22731
Angewandte
Research Articles Chemie

of the positively charged and zwitterionic lipids. For a fixed Previous HD-VSFG studies by us and others have shown
concentration of monomeric aS, the aggregation increases unique orientation of water at different lipid-water interfaces
linearly with the concentration of negatively charged lip- according to the charge of the lipid head groups.[28–33]
id.[9, 14] However, it is intriguing to observe that a negatively In this article, we clarified the lipid head group specificity
charged lipid surface promotes the aggregation of a net towards the interaction with aS and the associated reorgan-
negatively charged (@9) protein which is supposed to be ization of interfacial water using surface tensiometry and
repelled from the membrane surface. broadband HD-VSFG measurements. The air–water and
Apart from lipid head groups, interfacial water is believed various lipid monolayer-water interfaces (e.g., cationic 1,2-
to be crucial for the aggregation process. Time-resolved dipalmitoyl-3-trimethylammonium propane chloride salt
fluorescence measurement showed slower solvation of water (DPTAP), anionic 1,2-dipalmitoyl-sn-glycero-3-phospho-rac-
near negatively charged surfactant head group which pro- 1-glycerol sodium salt (DPPG), 1, 2-dipalmitoyl-sn-glycero-3-
motes fibrillization of disordered protein.[15, 16] MD simulation phosphoserine sodium salt (DPPS) as well as zwitterionic 1,2-
combined with time-resolved fluorescence suggests longer dipalmitoyl-sn-glycero-3-phosphocholine (DPPC); (Fig-
trapping of interfacial water (slower solvation) near the ure 1 b) were measured in the presence and absence of aS
amyloidogenic region that triggers pathological conversion of in the respective aqueous phases. The Imc(2) spectra (OH and
aS into amyloid form.[17] “Electron” and “atomic force” CH stretch regions) display significant effect of aS at the air–
microscopy measurements also reveal that the air–water water and charged lipid-water interfaces: the “H-down”
interface affects the pre-nucleation and aggregation of aS.[18] oriented water becomes “H-up” in presence of aS (20 mM) at
Undoubtedly, the aggregation of aS at membrane-water air/water interface, signifying the accumulation of aS at the
interface is a complex interplay of lipid-protein, lipid-water, air–water interface that makes the interface negatively
protein-water, and water-water interactions. Conventional charged. At the lipid monolayer-water interface, aS is
spectroscopic studies, such as circular dichroism (CD),[9] electrostatically attracted to the charged lipid head groups;
molecular probe-based fluorescence lifetime and microsco- however, it hardly interacts with the zwitterionic lipid (net
py,[18, 19] provide information about the changing rate of aS neutral) head group. Specifically, at the cationic DPTAP-
aggregation at membrane surface, but the role of interfacial water interface, the H-down orientation of interfacial water
water and the molecular origin of lipid head group-specificity decreases in the presence of aS (20–50 mM) due to screening
towards the accelerated aggregation are not well understood. of the lipid positive charge. In the case of anionic DPPG- and
Interface-selective spectroscopy such as vibrational sum DPPS-water interfaces, aS (20–50 mM) moderately decreases
frequency generation (VSFG) can provide molecular level the net H-up orientation of interfacial water, suggesting the
insight into the restructuring of interfacial water and the interaction of positively charged N-terminal of aS with
associated lipid-protein interaction during pre-nucleation of negatively charged lipid head groups. Thus, despite being
aS at model membrane-water interfaces. Understanding the net negative (@9), aS attractively interacts with anionic lipids
pre-nucleation at membrane-interface is important as the via its N-terminal at the membrane-water interface.
monomeric aS interacts with the lipid head group during this
stage which leads to the primary and secondary nucleation for
the fibrillation process.[9, 11] VSFG spectroscopy has been Results and Discussion
extensively used to unravel water structure and protein
conformational changes at aqueous interfaces.[20–23] Further- Figure 2 a shows the Imc(2) spectra (2660–3660 cm@1) of
more, heterodyne detection of VSFG signal (heterodyne- the neat air–water and air–water-aS (20 and 50 mM) inter-
detected VSFG), enabled us to directly measure the preferred faces. The Imc(2) spectrum of the air–water interface (black
orientation of interfacial molecules through the sign of Imc(2), dashed curve) shows a broad negative band, centered around
which is particularly advantageous in pinpointing the specific 3450 cm@1, and a positive band above & 3580 cm@1. In agree-
interaction between charged groups at the interface.[24–27] ment with previous reports,[34–36] the negative Imc(2) band is

Figure 2. a) Imc(2) spectra of the neat air–water and air–water-aS interfaces (surface pressure, p = 20 : 3 mN m@1); Inset: expanded view of the
OH stretch region. b) Imc(2) spectra of cationic lipid (DPTAP)-water and DPTAP-water-aS interfaces (p = 35 : 3 mN m@1).

22732 www.angewandte.org T 2020 Wiley-VCH GmbH Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737
Angewandte
Research Articles Chemie

due to “H-down” orientated interfacial water; whereas the present at the 20 mM aS solution. However, the positive Imc(2)
positive band above 3580 cm@1 whose maximum is signal in the OH stretch region (3000–3500 cm@1) decreases,
& 3700 cm@1 (not clearly visible in the present measurement though marginally. It indicates reduced orientational prefer-
due to insufficient IR power in that region) is due to “H-up” ence of interfacial water at higher concentration of aS
orientated dangling OH of the topmost water. It is important (50 mM). The decreased water signal suggests a possible
to note that previous phase-sensitive/heterodyne-detected conformational change of aS at the air–water interface such
VSFG measurements[20, 27, 37, 38] showed a positive Imc(2) band that the relative influence of the positively charged hydro-
around 3200 cm@1 and was assigned to the “ice-like” or phobic N-terminal of aS increases on interfacial water. This
“strongly H-bonded water pair”. The positive band however proposition agrees with recent electron- and atomic force
is not so prominent in the carefully measured recent Imc(2) microscopic measurements that suggested a conducive role of
spectra,[35, 36, 39–41] which is in conformity with the present data the aqueous interface towards the primary and secondary
(black-dashed curve in Figure 2 a ) measured using high nucleation and aggregation of aS.[18] The positive band around
resolution displacement sensor (see the Supporting Informa- 3600 cm@1 is clearly visible even for 50 mM aS, pointing to the
tion) as well as our previously reported spectra that were existence of H-up oriented weakly interacting OH at the air–
phase-corrected by internal referencing.[42–44] With the addi- water-aS interface. Given the hydrophobic nature of the N-
tion of 20 mM aS in water, the surface pressure (p = surface terminal, the existence of such weakly interacting water is
tension of neat water (72.8 mN m@1)—surface tension of aS quite likely at the air–water–aS interface.[28, 46] We note that
solution) becomes & 20 mN m@1, indicating adsorption of aS the peak position of the OH stretch band appears around
on water surface. In the HD-VSFG measurement too, the CH 3200 cm@1 in the presence of aS, which is red-shifted from that
stretch region (< 3000 cm@1) which was featureless for the of the neat air–water interface ( & 3450 cm@1). The apparent
neat air–water interface, shows sharp negative bands around red-shift may not be solely due to stronger H-bonding of
2880 and 2940 cm@1 (yellow curve) corresponding to the C@H interfacial water; the spectral deformation due to vibrational
stretch of aS, confirming the presence of aS at the air–water coupling of interfacial water[47] and the reversal of the sign of
interface. The OH stretch band ( & 3000–3580 cm@1) becomes Imc(2) in presence of aS may also contribute to the apparent
positive in sign due to interfacial aS (yellow curve, expanded maximum around 3200 cm@1.
view of the OH stretch is shown in the inset, Figure 2 a). The Next, we measure the Imc(2) spectra (Figure 2 b) of the
Positive sign implies that the interfacial water which was H- cationic DPTAP-water interface in the presence and absence
down oriented (negative Imc(2)) at the neat air–water inter- of aS (20 and 50 mM). The Imc(2) spectrum of the DPTAP-
face, adopts H-up orientation in presence of aS. At pH 7.4, water interface (solid black curve) shows a negative double-
monomeric aS contains both positively and negatively peak feature in the CH stretch region corresponding to the
charged domains (Figure 1 a) however, the net charge is terminal methyl symmetric stretch (CH3-SS, 2880 cm@1) and
negative (@9). Thus, it is the net charge of aS that makes the Fermi resonance with its bend overtone (CH3-FR,
interface negatively charged and preferentially orients the 2940 cm@1).[48] The CH3-SS band appears with a negative sign
interfacial water as H-up (positive Imc(2)), though it is not due to “CH3-up” orientation (i.e., methyl hydrogens are
obvious whether the net charge on the interfacial aS is @9 or pointed towards the air or away from the aqueous phase) of
less presumably due to the modified protonation Ð depro- DPTAP alkyl chains on the water surface.[27] A weak shoulder
tonation equilibrium at the interface. A cursory glance to the band around 2850 cm@1 corresponds to the methylene sym-
OH stretch band (at air–water-aS) reveals that its amplitude metric stretch (CH2-SS) of the alkyl chains. Although, the
is significantly lower than that of charged lipid-water inter- number density of -CH2- is higher than that of -CH3, the CH2-
faces (compare with the solid curves in Figure 2 a and 3 a). It SS signal is much weaker than CH3-SS due to local inversion
can be concluded that the adsorption of aS makes the symmetry in the well-packed alkyl chains of DPTAP mono-
interface only weakly negatively charged, which is either due layer.[49] In the high frequency region of the CH stretch,
to partial protonation of aS at the interface or due to the a positive band is observed around 2965 cm@1, corresponding
immersion of the negatively charged strongly hydrated to the CH3-AS of the terminal methyl of the lipid. The C@H
carboxyl groups deep into the aqueous phase such that they stretch signal of DPTAP remains almost invariant with 20 mM
contribute negligibly to the net orientation of water at the aS, but decreases slightly on increasing the concentration of
interface. The dip feature observed around 3280 cm@1 (yellow aS to 50 mM.
curve, Figure 2 a) is presumably due to the N-H stretch of The OH stretch signal at the DPTAP-water interface is
protein as observed in the chiral-VSFG study.[13] In the region significantly higher than the neat air–water interface (com-
above 3580 cm@1, the positive signal though noisy is distinctly pare solid- and dashed-black curves in Figure 2 b) due to
observable in the presence of aS (20 mM). increased preferential H-down orientation of water at the
With the increasing bulk concentration of aS (50 mM; blue cationic lipid interface following the principle of electro-
curve, Figure 2 a), the Imc(2) signal increases in the CH stretch statics. The Imc(2) spectrum shows a strong negative band in
region, suggesting a higher number density of aS at the 3000–3570 cm@1 regions, followed by a weak positive band
interface. To avoid aggregation in bulk water, we did not above 3570 cm@1. The intense negative band is due to the
increase aS concentration beyond 50 mM.[9, 45] Since aS is net large population of H-down oriented water in the electric
negative (@9; pH 7.4), its increased number density is double layer (EDL) below the positively charged lipid
expected to enhance negative charge density at the interface, monolayer; while the weak positive band is due to weakly
augmenting the H-up orientation of interfacial water already interacting H-up oriented water in the glycerol region of the

Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737 T 2020 Wiley-VCH GmbH www.angewandte.org 22733
Angewandte
Research Articles Chemie

lipid.[28, 32] It is worth noting that the opposite orientations of marginally changes the amplitudes of the CH stretch bands.
CH3-SS and OH-SS (methyl-up vs. H-down) corresponding to In the OH stretch region ( & 3000–3650 cm@1), the Imc(2)
the same negative sign of Imc(2) is due to the opposite signs of spectrum is positive in sign and high in amplitude due to
the hyperpolarizability (b) of CH3-SS and OH-SS vibrational preferred H-up orientation of water at the negatively charged
modes.[27] DPPG interface and the associated EDL region.[29, 32, 33]
With the addition of 20 mM aS, the negative Imc(2) signal Presence of aS (20 mM) slightly decreases the positive Imc(2)
(OH stretch; red curve) decreases and the central frequency signal (compare red and black curves, Figure 3 a). On
of the band shifts towards higher frequency. This result clearly increasing the concentration of aS (50 mM), the positive OH
suggests that the net negative aS screens the positive electric stretch signal decreases further, but not as significant as that
field of the DPTAP monolayer in the interfacial region such of the positively charged DPTAP interface. The decreased
that thickness of the EDL containing the preferentially H- positive Imc(2) signal is a manifestation of the reduced
down oriented water decreases. On increasing the concen- negative surface electric field at the DPPG-monolayer in
tration of aS to 50 mM, the negative OH stretch signal the presence of aS, presumably via the attractive interaction
decreases to such an extent that it becomes comparable to between the anionic lipid headgroup and the positively
that of the pristine air–water interface (compare solid-green charged N-terminal of aS. In the case of another negatively
and dashed-black curves, Figure 2 b). With the higher con- charged lipid, DPPS, similar spectral changes are observed in
centration of aS (50 mM) the observed spectral changes at the the presence of 20 and 50 mM aS (Figure 3 b). DPPS has the
DPTAP-water interface are instantaneous, while those at the same net charge (@1) as that on DPPG, but the total number
lower concentration of aS (20 mM) show modest time- of charged moieties in its head group are different (PO4@ ,
dependence (see Figure S4 in the Supporting Information). COO@ and NH3+; Figure 1 b). From the perspective of
The Imc(2) spectra shown in Figure 2 and subsequently are interfacial water, similar spectral changes at the DPPG and
equilibrated spectra recorded after a waiting time of several DPPS interfaces confirm that aS interacts with net negative
hours. The spectral changes in Figure 2 b clearly show that the lipids with equal efficiency even when the lipids have
negatively charged aS is electrostatically attracted towards a different distribution of charge moieties in their head
the positively charged head group of DPTAP such that the groups. This observation confirms that the lipid-aS interac-
water in the interfacial region does not experience any net tion is electrostatic in nature.
electric field (i.e., similar to the uncharged air–water inter- For further insight into the interaction, we performed MD
face) at 50 mM bulk concentration of aS. In other words, the simulation of DPPG-water and DPPG-water-aS systems (see
aS completely screens the positive charge of the lipid head the Supporting Information). The positional probability
group and the corresponding Imc(2) signal in 3100–3650 cm@1 calculation (Figure S7b) suggests an overlapping region
region selectively represents the response of the interfacial between the phosphate (PO4@) group of DPPG and aS, in
water free from the bulk-like oriented water in the EDL[50, 51] particular, with the positively charged lysine (LYS) at the N-
at the DPTAP-water-aS interface. The blue-shift of the terminal of aS. The radial distribution functions (RDFs)
central frequency of the Imc(2) spectra (OH stretch) suggests between PO4@ of DPPG and LYS (or aS) shows a strong
that the vibrational coupling and/or H-bonding of interfacial interaction between LYS and PO4@ (Figure S7c), reinforcing
water decrease in presence of aS at the cationic DPTAP- the notion that the positively charged LYS at the N-terminal
water interface. of aS is electrostatically attracted towards the DPPG head-
The Imc(2) spectrum of the negatively charged DPPG- group. It is important to mention that we did not find any
water interface (black curve Figure 3 a) shows two negative significant overlap between the hydrocarbon chain of DPPG
and one positive band in the CH stretch region, corresponding and aS, suggesting that the interaction is localized within the
to the CH3-SS, CH3-FR, and CH3-AS bands, similar to that of aqueous interfacial region without a noticeable effect on the
the DPTAP-water interface (Figure 2 b). Presence of aS lipid hydrophobic region. Orientational distributions of water

Figure 3. a) Imc(2) spectra of anionic DPPG-water and DPPG-water-aS interfaces (p = 35 : 3 mN m@1). b) Imc(2) spectra of net negative DPPS-
water and DPPS-water-aS interfaces (p = 35 : 3 mN m@1). The spectrum of the neat air–water interface (dashed curve) is shown in each panel for
reference.
22734 www.angewandte.org T 2020 Wiley-VCH GmbH Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737
Angewandte
Research Articles Chemie

depict H-up orientation of water at the DPPG-water at Accordingly, the H-up oriented interfacial water that is, the
DPPG-water-aS systems (Figure S8), which is in line with the positive Imc(2) signal (OH stretch) is expected to remain
experimental results. However, unlike DPPG, water near to unaffected (if not increased) in the presence of aS at the
the aS have both the “H-up” and “H-down” orientations, DPPG- and DPPS-water interfaces. However, the experi-
which could be due to the contribution of different local mental results (Figure 3 a and b) show that the positive Imc(2)
amino acids of aS. signal (OH stretch) decrease in presence of aS at these
In addition to the charged lipids, we measured the interfaces, revealing the screening of the lipid head group
interaction of aS with zwitterionic DPPC lipid (Figure 4). negative charge. We propose that the positively charged N-
Imc(2) spectrum of the DPPC-water interface (in absence of terminal of aS preferentially interacts with the negatively
aS) shows typical CH stretch bands below 3000 cm@1, charged head group of DPPG (or DPPS) such that the net
followed by a positive band around 3200 cm@1 and a dip- negative aS exhibits an attractive interaction with the
feature around 3450 cm@1. The 3200 cm@1 band is mainly due negatively charged membrane headgroups and screens the
to anionic phosphate associated water while the dip-feature negative electric field to some extent in the interfacial region.
around 3450 cm@1 is predominantly due to the cationic Nevertheless, the significant positive OH stretch signal at the
choline associated water.[30] Presence of aS (up to 50 mM) DPPG and DPPS interfaces, even in presence of 50 mM aS,
hardly affects the Imc(2) spectrum at the DPPC-water inter- indicates that the surface field is still strong enough to
face, which is a manifestation of negligible interaction of aS preferentially orient the interfacial water as H-up. Currently,
with the zwitterionic DPPC. This observation in combination it is not clear whether the negative electric field originates
with the results at the charged lipid interfaces further suggests from the inherent negative charge of the lipid headgroup or
that the hydrophobic alkyl chains of the lipids do not directly the interfacial aS (net negative) interacting with the lipid
interact with aS; it is mainly the lipid headgroups that interact headgroup via its positive N-terminal. Irrespective of the
electrostatically with the charged domains of aS. This origin of the net negative charge, it is confirmed that aS
observation qualitatively agrees with the less pronounced interacts with anionic lipid monolayer via electrostatic
aggregation of aS at zwitterionic lipid vesicles than that of the interaction, though from the perspective of the interfacial
negatively charged lipids.[8, 9] Since, many laboratories, instead water the screening effect of aS on the negative surface field
of Imc(2), report the SFG-intensity i.e., j c(2) j 2-spectra, we (DPPG and DPPS interface) is much weaker than that on the
showed the deduced j c(2) j 2-spectra corresponding to Imc(2) positive surface field (DPTAP interface). This observation is
spectra of Figure 2 b, 3a and 4, in the Supporting Information in conformity with the findings of several non-interface-
(Figure S5). selective spectroscopic investigations. For instance, the fluo-
In summary, our results reveal that aS effectively interacts rescence energy transfer study of aS in negatively charged
with those lipids that have charged head groups, and the SDS micelles suggests compaction of N-terminal and expan-
interaction is primarily electrostatic in nature. The net sion of C-terminal regions, leading to the conformational
negative aS (@9 in bulk water, pH 7.4) is attracted to the change of aS.[52] NMR[53] and EPR[54, 55] studies of aS with
cationic lipid headgroups, which drastically screens the synthetic lipid vesicles and detergent micelles also pointed out
positive charge and reduces the thickness of the EDL in the that the C-terminal of aS remains free whereas the N-
vicinity of the corresponding membrane surface. However, terminal binds to the vesicle and micelle surfaces, which
the most intriguing observation is the screening of the lipid triggers fibrillation. Fibril monomer equilibration study of aS
negative charge (DPPG and DPPS) by the net negative aS. with the varying charge on the C-terminal suggests that
Anionic lipids are expected to repel the negatively charged aS fibrillation is favored for the C-terminal of lower negative
(Figure 1 a) such that the interface will be depleted of aS. charge.[56, 57] Similar observation of decrease in the aggrega-
tion of aS by its C-terminus is also reported in microscopy,
circular dichroism and FT-IR studies.[58, 59]

Conclusion

We probed the interaction of aS with cationic, anionic,


and zwitterionic lipid monolayers on the water surface by
using interface-selective heterodyne-detected vibrational
sum-frequency generation (HD-VSFG) spectroscopy in the
OH and CH stretch regions. The Imc(2) spectra of the air–
water-aS interface show that the aS adsorbs at the air–water
interface and invert the net orientation of interfacial water to
H-up (i.e., water Hs are pointed towards the air and away
from the aqueous phase), corresponding to a weakly neg-
atively charged aqueous interface. The changes in the Imc(2)
Figure 4. Imc(2) spectra of the zwitterionic DPPC-water (black) and
DPPC-water-aS interfaces (red and green; p = 35 : 3 mN m@1). The spectra exposes the strong interaction of aS with the
spectrum of the neat air–water interface (dashed curve) is shown as positively charged lipid (DPTAP)-water interface, making
reference. the interface net neutral from the perspective of the

Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737 T 2020 Wiley-VCH GmbH www.angewandte.org 22735
Angewandte
Research Articles Chemie

interfacial water. The net negative aS also interacts with the [13] E. C. Y. Yan, Z. Wang, L. Fu, J. Phys. Chem. B 2015, 119, 2769 –
anionic lipid (DPPG and DPPS) headgroups presumably via 2785.
the positively charged N-terminal of aS; though the effective [14] K. Pirc, V. Hodnik, N. P. Ulrih, G. Anderluh, Croat. Chem. Acta
2016, 89, 255 – 260.
screening of the interfacial electric field is much weaker than
[15] S. Arya, S. Mukhopadhyay, J. Phys. Chem. B 2014, 118, 9191 –
that at the cationic lipid-water interface. In the case of
9198.
zwitterionic lipid (DPPC), aS does not produce noticeable [16] D. Thirumalai, G. Reddy, J. E. Straub, Acc. Chem. Res. 2012, 45,
changes in the CH and OH stretch regions, suggesting no 83 – 92.
preference of aS for the zwitterionic lipid-water interface. [17] S. Arya, A. K. Singh, T. Khan, M. Bhattacharya, A. Datta, S.
Altogether, the Imc(2) spectral changes suggest that aS Mukhopadhyay, J. Phys. Chem. Lett. 2016, 7, 4105 – 4110.
interacts with the charged lipid head groups at the aqueous [18] S. Campioni, G. Carret, S. Jordens, L. Nicoud, R. Mezzenga, R.
interface and the hydrophobic alkyl chains of the lipids Riek, J. Am. Chem. Soc. 2014, 136, 2866 – 2875.
remain largely unperturbed. The aS-lipid interaction and the [19] S. Nath, J. Meuvis, J. Hendrix, S. A. Carl, Y. Engelborghs,
Biophys. J. 2010, 98, 1302 – 1311.
associated change in water structure and orientation may play
[20] C.-S. Tian, Y. R. Shen, J. Am. Chem. Soc. 2009, 131, 2790 – 2791.
a crucial role for the fibrillation of aS at the membrane [21] L. Fu, J. Liu, E. C. Y. Yan, J. Am. Chem. Soc. 2011, 133, 8094 –
surface. 8097.
[22] M. F. M. Engel, C. C. vandenAkker, M. Schleeger, K. P. Velikov,
G. H. Koenderink, M. Bonn, J. Am. Chem. Soc. 2012, 134,
Acknowledgements 14781 – 14788.
[23] P. A. Covert, D. K. Hore, Annu. Rev. Phys. Chem. 2016, 67, 233 –
PCS acknowledges DST-SERB-India [Grant Number: CRG/ 257.
[24] I. V. Stiopkin, H. D. Jayathilake, A. N. Bordenyuk, A. V. Bend-
2019/001389] for financial support. JAM gratefully acknowl-
erskii, J. Am. Chem. Soc. 2008, 130, 2271 – 2275.
edges the support and encouragement from Dr. Awadhesh [25] S. Nihonyanagi, J. A. Mondal, S. Yamaguchi, T. Tahara, Annu.
Kumar, Head, RPCD and Dr. A. K. Tyagi, Associate Rev. Phys. Chem. 2013, 64, 579 – 603.
Director, Chemistry Group, BARC. BB thanks Dr. Sampali [26] Y. R. Shen, Annu. Rev. Phys. Chem. 2013, 64, 129 – 150.
Banerjee and Dr. Prosenjit SenQs lab of IACS for help in [27] S. Nihonyanagi, S. Yamaguchi, T. Tahara, J. Chem. Phys. 2009,
protein purification. BB and SR thank CSIR and HBNI, 130, 204704.
BARC for their respective fellowship. [28] Y. Nojima, Y. Suzuki, S. Yamaguchi, J. Phys. Chem. C 2017, 121,
2173 – 2180.
[29] P. C. Singh, K.-i. Inoue, S. Nihonyanagi, S. Yamaguchi, T. Tahara,
Angew. Chem. Int. Ed. 2016, 55, 10621 – 10625; Angew. Chem.
Conflict of interest 2016, 128, 10779 – 10783.
[30] J. A. Mondal, S. Nihonyanagi, S. Yamaguchi, T. Tahara, J. Am.
The authors declare no conflict of interest. Chem. Soc. 2012, 134, 7842 – 7850.
[31] S. Nihonyanagi, S. Yamaguchi, T. Tahara, J. Am. Chem. Soc.
Keywords: electrostatic interactions · lipids · interfaces · 2010, 132, 6867 – 6869.
membranes · Parkinson’s disease [32] J. A. Mondal, S. Nihonyanagi, S. Yamaguchi, T. Tahara, J. Am.
Chem. Soc. 2010, 132, 10656 – 10657.
[33] X. Chen, W. Hua, Z. Huang, H. C. Allen, J. Am. Chem. Soc.
2010, 132, 11336 – 11342.
[34] S. Sun, R. Liang, X. Xu, H. Zhu, Y. R. Shen, C. Tian, J. Chem.
[1] C. M. Dobson, Nature 2003, 426, 884 – 890. Phys. 2016, 144, 244711.
[2] C. M. Dobson, Nature 2002, 418, 729 – 730.
[35] S. Yamaguchi, J. Chem. Phys. 2015, 143, 034202.
[3] M. Goedert, Nat. Rev. Neurosci. 2001, 2, 492 – 501.
[36] S. Nihonyanagi, R. Kusaka, K.-i. Inoue, A. Adhikari, S.
[4] M. G. Spillantini, R. A. Crowther, R. Jakes, M. Hasegawa, M.
Yamaguchi, T. Tahara, J. Chem. Phys. 2015, 143, 124707.
Goedert, Proc. Natl. Acad. Sci. USA 1998, 95, 6469 – 6473.
[37] N. Ji, V. Ostroverkhov, C. S. Tian, Y. R. Shen, Phys. Rev. Lett.
[5] M. Grazia Spillantini, R. A. Crowther, R. Jakes, N. J. Cairns,
2008, 100, 096102.
P. L. Lantos, M. Goedert, Neurosci. Lett. 1998, 251, 205 – 208.
[38] S. Nihonyanagi, T. Ishiyama, T.-k. Lee, S. Yamaguchi, M. Bonn,
[6] C. Del Mar, E. A. Greenbaum, L. Mayne, S. W. Englander, V. L.
A. Morita, T. Tahara, J. Am. Chem. Soc. 2011, 133, 16875 – 16880.
Woods, Proc. Natl. Acad. Sci. USA 2005, 102, 15477.
[7] B. G. Wilhelm, S. Mandad, S. Truckenbrodt, K. Kroehnert, C. [39] X. Xu, Y. R. Shen, C. Tian, J. Chem. Phys. 2019, 150, 144701.
Schaefer, B. Rammner, S. J. Koo, G. A. Classen, M. Krauss, V. [40] M. Ahmed, Y. Nojima, S. Nihonyanagi, S. Yamaguchi, T. Tahara,
Haucke, H. Urlaub, S. O. Rizzoli, Science 2014, 344, 1023 – 1028. J. Chem. Phys. 2020, 152, 237101.
[8] C. Galvagnion, J. W. P. Brown, M. M. Ouberai, P. Flagmeier, M. [41] X. Xu, Y. R. Shen, C. Tian, J. Chem. Phys. 2020, 152, 237102.
Vendruscolo, A. K. Buell, E. Sparr, C. M. Dobson, Proc. Natl. [42] N. Ghosh, A. K. Singh, J. A. Mondal, J. Phys. Chem. C 2016, 120,
Acad. Sci. USA 2016, 113, 7065. 23596 – 23603.
[9] C. Galvagnion, A. K. Buell, G. Meisl, T. C. T. Michaels, M. [43] J. A. Mondal, J. Phys. Chem. Lett. 2016, 7, 1704 – 1708.
Vendruscolo, T. P. J. Knowles, C. M. Dobson, Nat. Chem. Biol. [44] J. A. Mondal, V. Namboodiri, P. Mathi, A. K. Singh, J. Phys.
2015, 11, 229 – 234. Chem. Lett. 2017, 8, 1637 – 1644.
[10] E. I. OQLeary, J. C. Lee, Biochim. Biophys. Acta Proteins [45] K. Bhasne, N. Jain, R. Karnawat, S. Arya, A. Majumdar, A.
Proteomics 2019, 1867, 483 – 491. Singh, S. Mukhopadhyay, J. Phys. Chem. B 2020, 124, 708 – 717.
[11] R. Gaspar, E. Sparr, G. Meisl, A. K. Buell, T. P. J. Knowles, L. [46] S. Roy, B. Biswas, J. A. Mondal, P. C. Singh, J. Phys. Chem. C
Young, C. F. Kaminski, S. Linse, Q. Rev. Biophys. 2017, 50, e6. 2018, 122, 26928 – 26933.
[12] L. Fu, Z. Wang, V. S. Batista, E. C. Y. Yan, J. Diabetes Res. 2016, [47] M. Sovago, R. K. Campen, G. W. H. Wurpel, M. Muller, H. J.
https://doi.org/10.1155/2016/7293063. Bakker, M. Bonn, Phys. Rev. Lett. 2008, 100, 173901.

22736 www.angewandte.org T 2020 Wiley-VCH GmbH Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737
Angewandte
Research Articles Chemie

[48] J. C. Conboy, M. C. Messmer, G. L. Richmond, J. Phys. Chem. B [56] Y. Izawa, H. Tateno, H. Kameda, K. Hirakawa, K. Hato, H. Yagi,
1997, 101, 6724 – 6733. K. Hongo, T. Mizobata, Y. Kawata, Brain Behav. 2012, 2, 595 –
[49] J. F. D. Liljeblad, V. Bulone, E. Tyrode, M. W. Rutland, C. M. 605.
Johnson, Biophys. J. 2010, 98, L50 – L52. [57] K. Levitan, D. Chereau, S. I. A. Cohen, T. P. J. Knowles, C. M.
[50] K. C. Jena, P. A. Covert, D. K. Hore, J. Phys. Chem. Lett. 2011, 2, Dobson, A. L. Fink, J. P. Anderson, J. M. Goldstein, G. L.
1056 – 1061. Millhauser, J. Mol. Biol. 2011, 411, 329 – 333.
[51] S.-h. Urashima, A. Myalitsin, S. Nihonyanagi, T. Tahara, J. Phys. [58] W. Hoyer, D. Cherny, V. Subramaniam, T. M. Jovin, Biochem-
Chem. Lett. 2018, 9, 4109 – 4114. istry 2004, 43, 16233 – 16242.
[52] J. C. Lee, R. Langen, P. A. Hummel, H. B. Gray, J. R. Winkler, [59] I. V. J. Murray, B. I. Giasson, S. M. Quinn, V. Koppaka, P. H.
Proc. Natl. Acad. Sci. USA 2004, 101, 16466 – 16471. Axelsen, H. Ischiropoulos, J. Q. Trojanowski, V. M. Y. Lee,
[53] D. Eliezer, E. Kutluay, R. Bussell, G. Browne, J. Mol. Biol. 2001, Biochemistry 2003, 42, 8530 – 8540.
307, 1061 – 1073.
[54] M. Robotta, C. Hintze, S. Schildknecht, N. Zijlstra, C. Jgngst, C.
Karreman, M. Huber, M. Leist, V. Subramaniam, M. Drescher, Manuscript received: August 15, 2020
Biochemistry 2012, 51, 3960 – 3962. Revised manuscript received: August 27, 2020
[55] C. C. Jao, A. Der-Sarkissian, J. Chen, R. Langen, Proc. Natl. Accepted manuscript online: August 31, 2020
Acad. Sci. USA 2004, 101, 8331 – 8336. Version of record online: October 7, 2020

Angew. Chem. Int. Ed. 2020, 59, 22731 – 22737 T 2020 Wiley-VCH GmbH www.angewandte.org 22737

You might also like