Chemistry and Physics of Lipids: Rafael P. Barroso, Luis G.M. Basso, Antonio J. Costa-Filho

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Chemistry and Physics of Lipids 186 (2015) 68–78

Contents lists available at ScienceDirect

Chemistry and Physics of Lipids


journal homepage: www.elsevier.com/locate/chemphyslip

Interactions of the antimalarial amodiaquine with lipid model


membranes
Rafael P. Barroso, Luis G.M. Basso, Antonio J. Costa-Filho *
Laboratório de Biofísica Molecular, Departamento de Física, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto. Av. Bandeirantes, 3900, 14040-901
Ribeirao Preto, SP, Brazil

A R T I C L E I N F O A B S T R A C T

Article history: A detailed molecular description of the mechanism of action of the antimalarial drug amodiaquine (AQ) is
Received 12 March 2014 still an open issue. To gain further insights on that, we studied the interactions of AQ with lipid model
Received in revised form 24 December 2014 membranes composed of dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylserine
Accepted 26 December 2014
(DPPS) by spin labeling electron spin resonance (ESR) and differential scanning calorimetry (DSC). Both
Available online 30 December 2014
techniques indicate a coexistence of an ordered DPPS-rich domain with a disordered DPPC-rich domain
in the binary DPPC/DPPS system. We found that AQ slightly lowered the melting transition temperatures
Keywords:
associated to both domains and significantly increased the enthalpy change of the whole DPPC/DPPS
Amodiaquine
DSC
phase transition. DSC and ESR data also suggest that AQ increases the number of DPPC molecules in the
ESR DPPC-rich domains. AQ also causes opposing ordering effects on different regions of the bilayer: while
Antimalarial the drug increases the ordering of the lipid acyl chains from carbon 7 to 16, it decreases the order
Spin label parameter of the lipid head group and of carbon 5. The gel phase was mostly affected by the presence of
AQ, suggesting that AQ is able to influence more organized lipid domains. Moreover, the effects of AQ and
cholesterol on lipid acyl chain ordering and mobility were compared at physiological temperature and, in
a general way, they are similar. Our results suggest that the quinoline ring of AQ is located completely
inside the lipid bilayers with its phenol ring and the tertiary amine directed towards the head group
region. The nonspecific interaction between AQ and DPPC/DPPS bilayers is a combination of electrostatic
and hydrophobic interactions.
ã 2014 Elsevier Ireland Ltd. All rights reserved.

1. Introduction low or absent, the therapeutic efficacy of artesunate–amodiaquine


is highly correlated with the efficacy of AQ alone (Sá et al., 2009).
Malaria is still one of the most infectious diseases affecting the The mechanism of action of 4-aminoquinolines is not
developing world. According to the World Health Organization completely understood, but it is believed that is related to
(WHO), 655,000 people die every year due to malaria, out of which hemoglobin degradation in the digestive vacuole of the parasite
86% are children under five years of age (WHO, 2011). One of the (Pussard and Verdier, 1994; O’Neill et al., 1998; Fitch, 2004).
major barriers on fighting malaria is the worldwide drug Hemoglobin is degraded by several enzymes of the parasite
resistance. Malaria parasites have developed resistance to all resulting in toxic heme metabolites. A process of detoxification
current drugs. To overcome drug resistance, WHO recommends the converts heme into an insoluble crystal called hemozoin or
use of combined therapies with two or more drugs, in particular, “malaria pigment”, which participates in the parasite life cycle
the combination of amodiaquine (AQ) and artesunate (WHO, (Sherman, 1977). A complex formed by 4-aminoquinolines with
2010). AQ is a 4-aminoquinoline antimalarial drug widely used in heme (ferriprotoporphyrin IX) inhibits conversion of heme to
the past and whose chemical structure is similar to that of hemozoin, thus killing the parasite. Although AQ forms complexes
chloroquine (Olliaro et al., 1996 O’Neill et al., 1998). Several with ferriprotoporphyrin IX (Blauer et al., 1993), its complete
countries chose AQ as the first-line drug in combination with mechanism of action, at a molecular level, remains unclear. In
artesunate (WHO, 2010). In areas where artesunate resistance is particular, there is not much information on how the drug
overcomes the physical barrier represented by the cell membrane.
Membrane lipids are associated with vital cell functions.
Besides being the structural blocks of cell membranes and the
* Corresponding author. Tel.: +55 163315 3665; fax: +55 16 33715381. sources of metabolic energy, lipids play a central role in the
E-mail address: ajcosta@ffclrp.usp.br (A.J. Costa-Filho). regulation of innumerous cellular processes (Escribá, 2006;

http://dx.doi.org/10.1016/j.chemphyslip.2014.12.003
0009-3084/ ã 2014 Elsevier Ireland Ltd. All rights reserved.
R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78 69

Bittman, 2004). Changes in the structure and lipid composition of and 0.5 mol% of spin label for ESR experiments, by drying under a
membranes seem to be related to the development of several stream of N2 and left under reduced pressure for 2 h to remove all
diseases (Lee, 2004). Although drug targets are usually proteins traces of organic solvent. The film was then hydrated with 10 mM
(Cohen, 2002; Overington et al., 2006), drugs can also bind to Hepes, at pH 7.4, to a lipid concentration of 20 mM, heated at 60  C
membrane lipids. Non-specific interactions of antiarrhythmics for 5 min, and vortexed. Part of the lipid dispersion was diluted
with myocardial membranes, for instance, are related to their with Hepes buffer to a final lipid concentration of 10 mM and
mode of action (Gourley, 1971). Amlodipine, an antihypertensive another part was diluted with a 2 mM drug buffered solution to
drug, also interacts non-specifically with membranes (Mason et al., final lipid and drug concentrations of 10 and 1 mM, respectively.
1991). In a previous work, the non-specific interaction of the For ESR measurements with cholesterol-containing samples, a
antimalarial drug primaquine with biological membranes was lipid film was formed from stock solutions of pure DPPC, DPPS,
suggested to represent an additional route in its overall mode of and cholesterol and the dried film was hydrated with Hepes to a
action and/or could be associated to its adverse effects (Basso et al., final concentration of 10 mM of DPPC and DPPS and 1 mM of
2011). The importance of the non-specific drug-membrane cholesterol.
interactions as a molecular event can be enhanced by a new
therapeutic approach called membrane-lipid therapy. It is based on 2.2.2. Electron spin resonance
the binding of drugs to lipids, thus modulating the structure of ESR measurements were performed with a JEOL FA-200
membranes (Escribá, 2006). To understand the overall mechanism X-band spectrometer (JEOL Ltd., Tokyo, Japan). Temperature was
of action of AQ, studies of AQ-membrane interaction are important controlled by a circulator bath that pumps fluid through an
since the molecular targets of this drug, either proteins or lipids aluminum radiator attached to the ESR resonant cavity as
from membranes, are not completely known yet. described by Alaouie and Smirnov (2006). The samples were
A combination of techniques is frequently used to study drug- placed into 1 mm inner diameter glass capillaries and then
membrane interactions (Seydel and Wiese, 2002). In particular, concentrated by centrifugation. The glass capillary was
electron spin resonance (ESR) and differential scanning calorime- accommodated within a quartz tube. The acquisition conditions
try (DSC), among several available techniques, have provided were: field modulation frequency, 100 kHz; field modulation
important contributions (Zhao et al., 2007; Budai et al., 2003; amplitude, from 1 to 2 G; sweep width, 100 G; microwave power,
Könczöl et al., 2005; Pentak et al., 2008; Hamilton et al., 1991; 10 mW. Data presented here are an average of duplicate experi-
Basso et al., 2011). Changes in membrane properties induced by ments and the uncertainties are the respective standard
drugs or other molecules are detected by lineshape alterations in deviations.
the ESR spectra or by changes in heat exchange in DSC experiments
(Duarte et al., 2008; Basso et al., 2011; Couto et al., 2011; Dyszy 2.2.3. Nonlinear least-squares (NLLS) simulations of ESR spectra
et al., 2013). NLLS analyses of the ESR spectra were performed using the
In this paper, interactions of the antimalarial drug AQ with lipid program Multicomponent (https://sites.google.com/site/alten-
membrane models were investigated by ESR and DSC. The bach/labview-programs/epr-programs/), a LabVIEW-based ver-
membrane models were constituted by an equimolar mixture of sion (Altenbach, 2014) of the traditional scheme of Freed et al. for
dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphospha- fitting ESR spectra of nitroxide spin probes (Schneider and Freed,
tidylserine (DPPS), which are representative lipids of the parasite 1989; Freed, 1976). The software calculates an ESR spectrum by
membrane (Hsiao et al., 1991). The effects of AQ on the melting solving the stochastic Liouville equation, allowing for the
behavior of DPPC/DPPS bilayers were monitored by DSC. Using spin quantification of the structural dynamics of spin-labeled
labels attached to different positions along the lipid acyl chains and biomolecules in terms of motional rates and orientational
to the head group, we investigated the alterations induced by AQ ordering imposed by the local environment around the nitroxide
on the ordering and dynamics of different regions of lipid bilayers radical.
by ESR. We also compared the effects caused by AQ with the effects The dynamics of spin-labeled lipids in lipid bilayers are better
caused by cholesterol on the structural dynamics of the binary described by an axially symmetric rotational diffusion tensor,
system at physiological temperature. A possible drug location in whose principal components, R// and R?, represent the rotational
the lipid bilayer was proposed. diffusion rates around axes parallel and perpendicular, respec-
tively, to the mean symmetry axis for the rotation, zR (Schneider
2. Material and methods and Freed, 1989). For n-PC spin labels (n = 5, 7, 10, 12, 14, and 16), zR
is defined along the symmetry axis of the lipid, i.e., the fully
2.1. Materials extended direction of the acyl chains (see Fig. S1B of the
Supporting information). For DOPTC, zR is defined to be collinear
The lipids 1,2-dipalmitoyl-sn-glycero-3-phosphocholine to both the N-C4 and the N—O bonds of the tempo-choline group
(DPPC), 1,2-dipalmitoyl-sn-glycero-3-phospho-L-serine (DPPS), (Ge and Freed, 1998; Ge and Freed, 2003), which makes it also
cholesterol, and the spin labels 1-palmitoyl-2-stearoyl- parallel to the magnetic xm axis, in which the magnetic frame is
(n-doxyl)-sn-glycero-3-phosphocholine (n-PCSL, where n = 5, 7, defined (xm points along the N—O bond, zm lies along the 2pz axis
10, 12, 14, and 16) and 1,2-dioleoyl-sn-glycero-3-phospho(tempo) of the nitrogen atom, and ym is perpendicular to both – see
choline (DOPTC) were obtained from Avanti Polar Lipids, Inc. Fig. S1A). We found in our simulations that the spectra are not very
(Alabaster, AL). The antimalarial drug AQ and the Hepes buffer sensitive to R//. Thus, to reduce the number of fitting parameters,
(4-(2-hydroxyethyl)-1-piperizineethanesulfonic) were purchased qffiffiffiffiffiffiffiffiffiffiffiffiffi
N  R== =R? was fixed to 1 and R ¼ R2? R== was varied instead. The
3
from Sigma Chem., Co. (St. Louis). All reagents were used without
further purification. mobility of the lipid acyl chains and the head groups was then
reported as R?.
2.2. Methods The calculated ESR spectra of the spin-labeled lipids are the
result of the so-called MOMD (microscopic order-macroscopic
2.2.1. Lipid dispersion preparation disorder) model that takes into account the tendency of the spin
A lipid film was formed from a chloroform: methanol (2:1) labels to become partially ordered with respect to a local director
solution of DPPC:DPPS (1:1) or DPPC:DPPS:cholesterol (10:10:1), zD (defined as the normal vector to the bilayer surface), but these
70 R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78
[(Fig._1)TD$IG]

(a) (b)

Fig. 1. DSC traces of samples with 10 mM DPPC/DPPS in Hepes buffer with (gray) and without (black) AQ. Inset: zoom of the calorimetric DSC peaks.

local directors are randomly oriented (macroscopically disor- 3. Results and discussions
dered) in the sample (Meirovitch et al., 1984). The microscopic
molecular ordering of the spin labels is characterized by the order 3.1. Calorimetric results show AQ-bilayer interaction
parameters S0 and S2, which are manifested through an orienting
potential on the lipid head groups and the acyl chains that restricts The effect of AQ on lipid membranes was investigated by
the range of orientations of the spin probes (Schneider and Freed, determining the thermal behavior of a DPPC/DPPS mixture. Fig. 1
1989; Budil et al., 1996). S0 is a measure of the inclination of zR with shows the DSC traces of DPPC/DPPS dispersions with and without
respect to the local director, zD, whereas the nonsymmetric order AQ. Upon heating, the AQ-free samples exhibited a nonideal lipid
parameter S2 gives the deviation from cylindrical symmetry of the mixing, as also reported by other authors (Sevcsik et al., 2008),
molecular alignment of zR relative to zD. The larger the S0, the characterized by a broad melting transition. Nonideal lipid mixing
larger is the alignment, and the more restricted is the spin label means that the lipids are not randomly distributed in the
motion. membrane but organized in clusters or domains. In this situation
The parameters involved in the simulations were: g-tensor (gxx, the bilayer displays lateral heterogeneity. This lateral phase
gyy, gzz) and hyperfine-tensor (Axx, Ayy, Azz) components, rotational separation was previously observed in DPPC/DPPS mixtures (Luna
diffusion rates (R?, R//), orienting potential coefficients (c20, c22), and McConnell, 1977; Gordon et al., 2006). DSC traces of AQ-free
and whenever needed, homogeneous lorentzian or Gaussian and AQ-containing samples presented two calorimetric peaks at
linewidths. Seed values of the magnetic tensors were obtained temperatures T1 and T2 (T1 < T2, see the inset of Fig. 1). We suggest
from Earle et al. (1994), but small variations in their values were that these peaks could be attributed to the melting of lipids in two
allowed during the fitting process. The strategy of the simulation different domains, probably DPPC-rich and DPPS-rich domains.
follows that of Basso et al. (2011). The thermograms also displayed very low enthalpic pre-transition
As for the uncertainties of the parameters from the NLLS peaks that are neither well defined nor reproducible and thus will
analyses, we found an estimated error of 0.01 for S0 and 5% not be discussed. The enthalpy change (DH) of the whole transition
for R? of n-PC spin labels, and of 0.006 for S0 and 3% for R? of and the values of T1 and T2 are presented in Table 1. As observed in
DOPTC. This means that we found good solutions for the Fig. 1, incorporation of AQ into the DPPC/DPPS lipid mixture
stochastic Liouville equation with different sets of c20, c22 (only increases the peak intensity at T1, slightly lowers both T1 and T2,
used in the simulations of DOPTC and 16-PCSL), and R?. In and significantly increases DH by (20  2)% on average (Table 1). In
general, good agreement between the experimental and a previous work, a decrease of the transition temperature was also
theoretical spectra is obtained whenever S0 and R? are both observed with primaquine, another aminoquinoline antimalarial
simultaneously increased within the uncertainties. Overall, the drug, in DMPC bilayers (Basso et al., 2011). However, a much
general trends observed in the ordering and dynamics of the smaller amount of primaquine (lipid/drug molar ratio 91:1)
spin labels for different sets of (S0, R?) values and for different was required to produce a similar effect than that caused by AQ
lipid sample preparations were all consistent, even though the (lipid/drug molar ratio 10:1). On the other hand, while the
differences in S0 and mostly in R? might be within the transition enthalpy of DMPC bilayers remained unchanged in the
uncertainty of the corresponding parameters.
Table 1
2.2.4. Differential scanning calorimetry Calorimetric parameters of the phase transition of the binary DPPC/DPPS
membranes obtained from the DSC traces.
DSC traces were obtained by heating the samples from 5 to
70  C at a scan rate of 20  C/h with the Microcalorimeter VP-DSC Sample DH (kcal/mol) T1 ( C) T2 ( C)
(Microcal, Northampton, MA) or the calorimeter Nano-DSC-II DPPC/DPPS 1 17.4 43.23 44.39
(Calorimetry Sciences Corporation, Lindon, Utah, USA). Baseline 2 18.9 43.19 44.40
subtractions and peak integrals were carried out using MicroCal DPPC/DPPS with AQ 1 21.3 43.00 44.20
2 22.2 43.00 44.12
Origin software.
R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78 71
[(Fig._2)TD$IG]

Fig. 2. Experimental (black) and best-fit (gray) ESR spectra of the head group spin label DOPTC and the acyl chain n-PCSL (n = 5, 7, 10, 12, 14, and 16) in DPPC/DPPS membranes
in the absence (left) and in the presence (right) of AQ. (a) One-component nonlinear least squares fits of DOPTC, 5-, 7-, 10-, and 12-PCSL ESR spectra. (b) Spectra of 14- and
16-PC spin labels were simulated with two components. The arrows show the second, more mobile, and less populated component. The spectra were acquired at 36  C.
72 R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78
[(Fig._3)TD$IG]
presence of primaquine, AQ induced a large increase in DH 1.0
of DPPC/DPPS membranes, suggesting different drug-lipid
interactions and thus different effects on membranes. DPPC/DPPS - comp 1
DPPC/DPPS - comp 2
Our DSC results show a clear interaction between AQ and the 0.8
lipid membranes, since AQ changes the shape of the thermograms, DOPTC
DPPC/DPPS/AQ - comp 1
decreases the melting temperatures and increases the transition DPPC/DPPS/AQ - comp 2

Order parameter
enthalpy of the binary system. These results could indicate a 0.6
penetration of AQ into the DPPC/DPPS bilayers, since molecules
that penetrate lipid bilayers usually lower the gel-to-fluid phase
transition (Jain and Wu, 1977). As mentioned before, the narrow 0.4
DSC peak at T1 (Fig. 1) can be due to the melting of either DPPC-rich
or DPPS-rich domains. Using infrared spectroscopy, Schultz et al.
showed the coexistence of DPPC-rich and DPPS-rich aggregates in a 0.2
binary DPPC/DPPS membranes and also showed that DPPC
molecules melt at a temperature (45.2  C) lower than that of
DPPS (47.8  C) in this lipid system (Schultz et al., 2009). Thus, we 0.0
propose that the peak at T1 in our DSC thermograms might be due HG 4 6 8 10 12 14 16
n-PCSL
to the melting of DPPC in DPPC-rich domains, whereas the peak at
T2 could be related to the melting of DPPS in DPPS-rich domains. Fig. 3. Order parameter profile obtained from non-linear least-squares simulations
Furthermore, lipids do not melt independently of each other but of DOPTC and n-PCSL (n = 5, 7, 10, 12, 14, and 16) 36- C ESR spectra in the absence
rather simultaneously in clusters of a certain amount of lipids (full circle) and in the presence (open triangle) of AQ. HG stands for head group.
(Heimburg, 2007). Thus, the melting of lipids is a cooperative
phenomenon and the cooperativity is proportional to the number ammonium (TMA) group of the DOPTC in the binary DPPC/DPPS
of lipids in a cluster or domain. The larger the cooperativity, the system is almost similar to that of the dipalmitoylphosphatidyl
narrower and more intense the transition peak. The intensity tempo (2,2,6,6,-tetramethyl-1-oxy) choline (DPPTC) in a pure DPPC
increase of the DSC peak at T1 caused by AQ binding to the binary dispersion in the gel phase (S0 = 0.50 and S2 = –0.22 at 25  C)
system is therefore likely due to an increase of the number of lipids (Ge and Freed, 1998) or to that of 4-O-(1,2-dipalmitoyl-sn-glycero-
in the domains melting at T1, that is, the size of the DPPC-rich 3-phospho) 4-hydroxy-2,2,6,6-tetramethyl-piperidine-1-oxy
domains increases. Since there is a direct correlation between (DPP-Tempo) in dioleoylphosphatidylcholine (DOPC) dispersions
membrane curvature and both enthalpy changes (Yokoyama et al. at –11  C (S0 = 0.516 and S2 = –0.238) in the presence of excess water
(2013)) and lateral phase separation (Duwe et al., 1990; Sackmann, (Ge and Freed, 2003). This means that the principal diffusion axis
2006), the aforementioned effects of AQ on the lipid mixture could (zR), which coincides with xm (parallel to the N—O bond – see
be explained by an increase in DPPC/DPPS membrane curvature. Section 2.2.3), is preferentially aligned along the local director of
the bilayer surface. The main difference in the orientation
3.2. ESR results suggest AQ penetration into the bilayer of the TMA group of DOPTC in DPPC/DPPS membranes compared
to that of the other spin labels (DPPTC in DPPC or DPP-Tempo
Changes in the order and dynamics of the membrane caused by in DOPC) is the lower preference of either xR or yR axes to be
AQ were investigated by means of ESR spin labeling technique at oriented perpendicular to director, due to the low value of S2
physiological temperature. The hydrophobic region of the bilayer (–0.084). To make this clear, we decomposed the spherical tensor
was monitored by using spin labels located at different depths components S0 and S2 in their Cartesian counterparts (Ge and
p
along the lipid acyl chain. Fig. 2a and b shows the experimental Freed, 2003): Szz = S0 = 0.518, Sxx = 1/2( 3/2 S2 – S0) = –0.310, and
1 p
and simulated ESR spectra of the head group spin label DOPTC Syy = – /2( 3/2 S2 + S0) = –0.207. As discussed before, the high
and the acyl chain spin labels n-PCSL (n = 5, 7, 10, 12, 14, and 16) positive value of Szz means a strong preference for zR to be aligned
acquired at 36  C in the absence and in the presence of AQ. along the normal to the bilayer, whereas negative and close values
Generally, the experimental and best-fit theoretical spectra were of Sxx and Syy indicate an almost equal preference for xR and yR
in very good agreement and considerable changes were observed diffusion axes to be oriented tangential to the bilayer. The addition
in all ESR spectra upon incorporation of AQ. Except for the 14- and of AQ to DPPC/DPPS samples changes the ordering and dynamics of
the 16-PCSL signals (Fig. 2b), all other ESR spectra were simulated DOPTC, which means that AQ binds to and interacts with the lipid
with only one component (one spin label population). The head group of the bilayers. Interestingly, AQ significantly decreases
two-component feature in the 14- and 16-PCSL spectra will be both the perpendicular rotational mobility R? (from 3.02 to
discussed below. 2.45  107 s1) and the order parameter S0 (from 0.518 to 0.452) of
Fig. 3 presents the order parameter (S0) profile obtained from the spin label. The latter result implies that lipid-AQ interaction
the NLLS simulations of the ESR spectra displayed in Fig. 2. Overall, increases the average angular amplitude of the wobbling motion of
NLLS fits show that AQ increases the ordering and slightly the nitroxide moiety (the larger S0, the smaller the amplitude). In
decreases the rotational mobility of the whole hydrophobic region terms of the Cartesian components, Sxx = –0.291, Syy = –0.160, and
of the lipid bilayers, except for the DOPTC and for 5-PCSL, in which Szz = 0.452, we can also infer that AQ binding to membranes
an opposite effect was observed: for example, AQ slightly induces zR to be much less oriented along the local director,
decreases S0 (from 0.711 to 0.692 – Table S2) and significantly whereas xR and yR axes retain their orientation tangential to the
increases R? (from 3.31 to 4.36  107 s1 – Table S2) for 5-PCSL. bilayer normal, but yR becomes less oriented.
Interesting features were reported by some of the spin labeled Another interesting feature observed in Fig. 2a is that the
lipids and will be discussed in more details. spectrum of 12-PCSL from the AQ-free sample presented a much
NLLS simulations of the ESR spectra of the head group spin label broader line than would be normally expected for this spin label.
DOPTC provided information on the dynamics and on the This clearly indicates strong spin–spin interactions that likely arise
orientation of the nitroxide radical at the bilayer surface. With a from cluster formation of 12-PCSL molecules in the DPPC/DPPS
positive value for S0 = 0.518 and a negative value for the nonaxial membranes. Partial segregation of spin labels was also noted, for
order parameter S2 = –0.084, the orientation of the trimethyl instance, by Fajer et al. for n-PCSL in the DPPC gel phase in a
R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78 73

saturation transfer ESR study (Fajer et al., 1992) and by Earle et al. result of two main effects: (1) changes in the equilibrium constant
in a rigid-limit 250 GHz ESR study of n-PCSL incorporated in DPPC between the DPPC-rich and DPPS-rich domains and/or (2) changes
and in palmitoyl-oleoyl-PC (POPC) lipid vesicles (Earle et al., 1994). in the partition coefficient of the spin labels between the two
In fact, the best least-square fits to 12-PCSL ESR spectra were regions. We explored these two possibilities as follows. Firstly, we
achieved by allowing for a Heisenberg exchange coupling in the defined the partition coefficient KP as the ratio of spin-label
simulations, which yielded a vHE = 4.9  107 s1. Interestingly, concentration in the ordered (Lo) and disordered (Ld) states as in
addition of AQ into the preformed 12-PCSL-clustered DPPC/DPPS Swamy et al. (2006):
lipid vesicles abolished the spectral broadening, thus suggesting
PðLo Þ ½mol%ofLd 
disruption of those clusters. Direct lipid-AQ interactions might KP ¼ x
PðLd Þ ½mol%ofLo 
promote a better miscibility of the 12-PCSL molecules in the binary
system.
where P(Ld) and P(Lo) are the populations of the spin labels
A distinct behavior was observed in the spectra of 14- and
incorporated in the disordered (Ld) and ordered (Lo) domains,
16-PCSL (Fig. 2b), which exhibited additional spectral features (see
respectively, obtained from the NLLS fits of the ESR spectra, and the
the low-field and high-field lines) characteristic of a second spin
second term on the right hand side of the equation represents the
population, indicating the coexistence of two types of domains in
percentage of lipids in the DPPC-rich (Ld) and DPPS-rich (Lo)
the 1:1 DPPC/DPPS lipid bilayers, which is in agreement with our
domains. KP would be unity if spin labels partition equally into
DSC results. A two-component system was also observed, for
each domain. If we further suppose that the relative populations
instance, by Basso et al. in the rippled phase of dimyristoyl-PC
(percentage of lipids) of the ordered and disordered domains are
(DMPC) lipid bilayers (Basso et al., 2011) and by Chiang et al. in a
characteristic of the binary system and thus do not change with the
ternary DPPC/dilauroyl-PC/cholesterol (DPPC/DLPC/Chol) lipid
addition of 0.5 mol% of the spin label (this is reasonable since the
mixture (Chiang et al., 2004). Our result shows that ESR is very
acyl chain of spin-labeled lipids are almost identical to that of DPPC
sensitive to detect the heterogeneity of different micro-environ-
and DPPS), the ratio K between KP values for 14- and 16-PCSL can be
ments in a binary system. Thus, an extra ESR signal with different
directly calculated as the ratio of the spin label populations
ordering and rotational dynamics was added to the fitting process.
obtained from the NLLS fits:
From the NLLS fits, we found that the most populated component
of the two spin labels presented the higher order parameter: S0 of K 14PC ½PðLo Þ=PðLd Þ14PC
P
0.367 for 14-PCSL (87% population) and of 0.343 for 16-PCSL (93% K¼ ¼
K 16PC
P
½PðLo Þ=PðLd Þ16PC
population), S0 values of the less ordered domain were 0.196 for
14-PCSL (13% population) and 0.117 for 16-PCSL (7% population).
As shown in Fig. S2 of the Supporting information, the theoretical Thus, K gives a good estimate of the relative partition coefficient
16-PCSL spectrum of the ordered domain (component 1) in the of both spin labels into the Ld and Lo phases. We found
binary system is very similar to the experimental spectrum of this that K = 0.504 for the AQ-free sample and K = 0.527 for the
same spin label in pure DPPS vesicles at the same temperature. AQ-containing sample, that is, the difference is within the
Similarly, the theoretical spectrum of the less ordered component uncertainty of the spin label populations obtained from NLLS
resembles, despite the difference in intensity, that of the 16-PCSL simulations, which is 3%. This result suggests that the partition-
incorporated in pure DPPC lipid bilayers at the same temperature ing of the spin labels in the ordered and disordered domains does
(Fig. S2). Thus, the coexistence of the ordered and disordered not depend on the AQ binding to membranes at 36  C (at least on
regions in the binary DPPC/DPPS lipid system is most likely due to the specific conditions of our experiments). Consequently, changes
the coexistence of DPPS-rich (ordered) and DPPC-rich (disordered) on spin label population upon AQ binding reflect changes on the
domains. equilibrium constant between DPPC-rich and DPPS-rich domains
The slight differences in the spin population values of the in the binary system. As mentioned before, our DSC experiments
ordered and disordered domains for 14- and 16-PCSL suggest that suggest an increase of the cooperativity of the DPPC-rich domains
these spin-labeled lipids have different partition coefficients KP upon AQ incorporation (see Section 3.1). Our ESR results also
between those domains. Since the mole fraction of 14-PCSL in the support this hypothesis. Binding of AQ to DPPC/DPPS membranes
less ordered state is higher than the population of 16-PCSL in this could shift the equilibrium constant between DPPC-rich and
same domain (0.13 and 0.07, respectively), 14-PCSL must partition DPPS-rich domains towards the DPPC-rich one. As a result, there
better in the disordered domains than 16-PCSL. In fact, a slighter may be an increase in the number of lipids in the DPPC-rich
preference of 14-PCSL for the disordered phase compared to clusters, which can contribute to the increased cooperativity of the
16-PCSL in a two-phase coexisting region of a ternary system has phase transition of this domain.
been reported (Swamy et al., 2006). Taken together, the ESR/NLLS results suggest that AQ inserts
Upon AQ addition to the lipid samples, changes in the lineshape into the DPPC/DPPS membranes and changes the molecular force
of the 14- and 16-PCSL are observed (Fig. 2b). The nonlinear least- field, i.e., the orienting potential, in the headgroup and in the acyl
squares fits show that binding of AQ to the DPPC/DPPS membranes chain regions. We suggest that both the phenol ring and the
increases the ordering (S0: from 0.367 to 0.413 for 14-PCSL and from protonated tertiary amine of AQ (at pH 7.4, AQ is mostly
0.343 to 0.379 for 16-PCSL) and decreases the perpendicular monoprotonated) are most likely located in the headgroup
rotational diffusion rate (R?: from 6.76 to 6.31 107 s1 for 14-PCSL region of the lipid bilayer. This can enhance the hydrogen bond
and from 2.09 to 1.78  108 s1 for 16-PCSL) of the DPPS-rich network of that region (due to additional hydrogen bonds between
domains (Table S2). The overall result is an increased packing of the AQ and the lipid head groups) and also decrease electrostatic
bilayer center of the DPPS-rich domain, whereas almost no change repulsion among PS in DPPS-rich domains. Direct AQ-head group
was observed on ordering and dynamics of the center of the interaction via the aforementioned groups also changes the
hydrophobic core of the DPPC-rich domains (Table S2). However, orientation of the TMA group of the DOPTC spin label, which
both 14- and 16-PCSL spectra of the AQ-containing samples show impacts the membrane orienting potential. Additionally, except for
an increase of the spin-labeled lipid population of the less the local environment around carbon 5, the increased S0 values of
ordered DPPC-rich domain (from 0.13 to 0.28, for 14-PCSL, and almost the entire acyl chain region indicate an increased lateral
from 0.07 to 0.17, for 16-PCSL). The redistribution of the spin label packing density of the lipids induced by AQ, which can enhance
molecules between the two domains upon AQ addition might be a van der Waals interactions between lipids. This increased bilayer
74 R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78

ordering is consistent with a higher transition enthalpy, since dynamics and order parameters obtained from the NLLS
more heat is required to break molecular interactions and melt the simulations. As before, the spectra of 14- and 16-PCSL were
lipids. The slight reduction of S0 with a remarkably increase of simulated with two components, with the greater contribution
rotational mobility around carbon 5 – as shown by NLLS (most populated) coming from the ordered, DPPS-rich domains,
simulations of 5-PCSL spectra – can be explained by the insertion for both spin labels (Table S2).
of the quinoline ring of AQ into the lipid matrix. With the quinoline Incorporation of 5 mol% of cholesterol into the binary
ring located into the hydrophobic core of the lipid bilayer, but close DPPC/DPPS system causes a rigidifying effect on the whole acyl
to the polar–apolar interface, and the phenol ring OH and the chain of the lipids (Table S2). Cholesterol also provokes a
tertiary amine directed towards the head group, a free volume depth-dependent effect on the perpendicular rotational diffusion
could be created around the very first carbon atoms of the lipid acyl rate of the lipids: a slight reduction of R? from carbon 5 to carbon
chains, which can explain the markedly increase of the rotational 7 and then a prominent increase of R? from carbon 12 to the end of
degree of freedom of 5-PCSL. Insertion of the quinoline structure the lipid acyl chain (Table S2). This depth-dependent effect of
into lipid matrix may also enhance van der Waals attractions cholesterol on the structural dynamics of lipid acyl chains in lipid
between lipids, which can contribute to increase the lateral bilayers is well known (Costa-Filho et al., 2003; Mainali et al.,
packing. In fact, insertion of both unprotonated and protonated 2013). Besides that, cholesterol causes no effect on the ordering of
forms of the quinoline ring of another 4-aminoquinoline the head groups (S0 = 0.52; S2 = –0.11), but slightly reduces the
antimalarial drug, chloroquine, into the apolar region of an anionic rotational mobility (from 3.16 to 3.02  107 s1), as monitored by
micelle has been also proposed (Santos et al., 2014; Usman and DOPTC spin label. Despite the small difference in R? (4.4%), this
Siddiq, 2013). variation is higher than the uncertainty calculated for this
To gain further insights on the hypothesis of AQ penetration parameter (3% – see Section 2.2.3). Also, the observed trend of
into the membrane, as strongly suggested by our results, we also reduced head group mobility in cholesterol-containing samples is
compared the effect of AQ on acyl chain order and mobility with obtained in different lipid preparations: there is a consistent
those provoked by cholesterol. It is well known that cholesterol is decrease of the ratio between the intensity of the low-field and the
located within the lipid bilayers (Huang, 1977; Ahmad and Mellors, center-field lines of the DOPTC ESR spectrum in the presence of
1978; Subczynski et al., 1992) and, as will be discussed further, cholesterol.
cholesterol is also soluble in PS bilayers. Changes in lipid order parameters can be used to infer how deep
Fig. 4 shows the ESR spectra of DOPTC and n-PCSL (n = 5, 7, 10, foreign molecules go inside the lipid bilayers (Biaggi et al., 1997). To
12, 14, and 16) in DPPC/DPPS (1:1) and DPPC/DPPS/Cholesterol gain further insights into the localization of AQ inside DPPC/DPPS
(1:1:0.1) model membranes and Table S2 summarizes the bilayers, we plotted the variation of S0 induced by AQ or cholesterol
[(Fig._4)TD$IG]

Fig. 4. Experimental (black) and best-fit (gray) of ESR spectra of the head group spin label DOPTC and the acyl chain n-PCSL (n = 5, 7, 10, 12, 14, and 16) in either DPPC/DPPS
(left) or DPPC/DPPS/Chol (right) membranes. Spectra of 14- and 16-PC spin labels were simulated with two components. All the spectra were acquired at 36  C.
R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78 75
[(Fig._5)TD$IG]
system (Fig 6a). Above the phase transition, there is no change in
the lineshape of the spectra due to addition of AQ. However,
cholesterol slightly decreases the ratio between the high-field and
the center-field lines, as indicated by the arrow in the figure, which
means that 5 mol% of Chol slightly decreases the mobility and/or
increases the ordering of the region monitored by 14-PCSL. On the
other hand, below the phase transition, both AQ and cholesterol
broaden the spectrum relative to the pure lipid mixture. This
means that both molecules increase the bilayer ordering, but the
effect caused by AQ is stronger than that of cholesterol, which
corroborates our initial hypothesis. The lineshape alterations in the
14-PCSL spectrum caused by incorporation of AQ into the inner and
outer monolayers (Fig. 6a) are even more pronounced than those
caused by incorporation of the drug only into the outer monolayer
(see Fig. 2b). The effect of cholesterol on the heat capacity of a
DPPC/DPPS/Chol lipid dispersion was also similar to the effect
caused by AQ, though there was no change in the transition
Fig. 5. Variation of the order parameter, S0, relative to the pure lipid bilayer as a enthalpy (Fig. 6b). Despite the low miscibility of cholesterol in
function of nitroxide position, n, of n-PCSL in DPPC/DPPS membranes in the phosphadityl serines (PS) bilayers reported by several authors
presence of AQ or cholesterol. Temperature: 36  C. (Bach and Wachtel, 2003), the effect of limiting amounts of
cholesterol on the thermotropic behavior of binary phospholipid
(Chol) relative to the pure lipid bilayer as a function of the labeled mixture indicated that anionic phospholipids exhibit greater
carbon atoms of the lipid acyl chain (Fig. 5). AQ and Chol affect the affinity for cholesterol than do the zwitterionic phospholipids
order parameter along the whole acyl chain with Chol effects PC and PE (Van Dijck, 1979). McMullen et al. (2000) also showed
decreasing from n = 7 down the chain, whereas AQ-induced that cholesterol exhibits greater miscibility in anionic PG than
alterations oscillate in the same region. Furthermore, the increase in uncharged glycolipid bilayers. A personal communication
in S0 at carbons 14 and 16 caused by AQ in the more ordered published in the McMullen et al. paper (McMullen et al., 2000)
domain is larger than that caused by cholesterol (component reports an up-to 67 mol% cholesterol solubility in 1-palmitoyl-
1 – Fig. 5). Unlike Chol, AQ also causes a significant perturbation of oleoyl-PS bilayers. Cholesterol incorporation, up to 15 mol%, in the
the headgroup organization, which indicates a distinct mode of positively charged DODAB bilayers was also reported (Benatti et al.,
organization of the drug in the bilayer. 2007). We used a cholesterol concentration of 5 mol% of the total
The previous experiments were carried out with AQ addition to lipid and, considering its reported higher affinity to PS compared
preformed liposomes (see Section 2), which means that only the to PC, we suggest that cholesterol partitions better in PS-rich
external part (the outer leaflet) of the bilayers was available to domains than in PC-rich ones.
interact with AQ. However, cholesterol was not incorporated into The incorporation of cholesterol in the bilayer is related to
the bilayers by the same procedure, but rather by preparing the hydrophobic effects between the cholesterol and the lipid bilayer.
liposome as a lipid film constituted of DPPC, DPPS, and cholesterol. Since AQ is a hydrophobic drug, and it is worth noting that
Thus, as both the inner and outer monolayers were available to its hydrophobicity has been linked to its antimalarial activity
cholesterol, the effect of AQ on the lipid ordering and mobility (Elslager et al., 1964), we expect that hydrophobic effects
might be even stronger than that caused by cholesterol. To test this between AQ and the bilayers could also take place. However,
hypothesis, we qualitatively studied the effect of AQ on the acyl besides the hydrophobic effect, electrostatic forces between AQ
chain ordering and dynamics when the drug was also allowed to and the negatively charged DPPC/DPPS bilayers may occur, since at
interact with the inner monolayers of the liposomes. In this the pH studied, about 98% of AQ was in the monoprotonated form
condition, we compared the effect of AQ and Chol on the lineshape (Naisbitt et al., 1997). To verify the role of the bilayer charge in the
of 14-PCSL spectra at temperatures above and below the phase interaction with AQ, we measured the effect of the same amount of
transition temperatures (above T2 and below T1) of the binary AQ on the ESR spectrum of 12-PCSL in DPPC/DPPS bilayers with
[(Fig._6)TD$IG]
(a) (b)

Fig. 6. (a) ESR spectra of 14-PCSL in DPPC/DPPS membranes in the absence and in the presence of AQ or Chol, in the gel phase (left) and fluid phase (right). (b) DSC traces of
10 mM DPPC/DPPS in Hepes buffer with (gray) and without (black) cholesterol.
76 R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78
[(Fig._7)TD$IG]
to interact with membranes, since the drug caused changes in
the thermotropic behavior of DPPS bilayers by decreasing the
transition enthalpy (Pajeva et al., 1996). Primaquine-membrane
interactions were shown by a combination of pressure perturba-
tion calorimetry, DSC and ESR (Basso et al., 2011). It was shown that
primaquine stabilizes the fluid phase of different lipid model
membranes, causing a local disorder on the acyl chains of the
lipids. This fluidizing effect seems to be due to the interaction of
primaquine with the lipid head groups, causing a less dense gel
phase. Interactions of mefloquine, quinacrine, chloroquine, and
quinine with DPPC bilayers were investigated by 2H and 31P NMR
(Zidovetzki et al., 1989). Although chloroquine did not seem to
affect the bilayer structure of DPPC, a weak chloroquine–DPPC
interaction could not be disregarded by the authors. Their results
also pointed that quinacrine interacts only with the surface of the
DPPC bilayers, whereas both quinine and mefloquine penetrate
into the membrane. A further investigation of DPPC–chloroquine
interactions using a fluorescent probe inserted into the bilayer
showed that chloroquine rigidified DPPC membranes (Ghosh et al.,
1995). Our results also show a rigidifying effect on the hydrophobic
core of the bilayers caused by AQ.
It is believed that the heme is the target of both AQ and
chloroquine, since these drugs prevent heme detoxification.
The parasite protects itself against the toxicity of heme by
polymerizing it to insoluble hemozoin. As weak bases, AQ and
chloroquine accumulate in the acidic environment of the parasite's
food vacuole and inhibit heme polymerization. After exiting the
food vacuole, the heme that is not polymerized is degraded by
glutathione. Ginsburg et al. found that AQ and chloroquine
Fig. 7. Ratio between the amplitudes of the low- and central-field lines (h+1/h0) of
competitively inhibit the degradation of heme by glutathione,
12-PCSL in 10 mM of DPPC and different amounts of DPPS in the absence (closed
square) and in the presence (open square) of AQ. Temperature: 36  C. allowing it to accumulate into the membranes (Ginsburg et al.,
1998). Heme, which is released as ferriprotoporphyrin IX (FPIX) in
the presence of oxygen (Jacob and Winterhalter, 1970), disrupts the
increasing amounts of DPPS (Fig. 7). Although variations can be membrane barrier properties and disturbs the ion homeostasis in
observed even in the absence of DPPS, which means that AQ also the cells, which kills the parasites. Famin et at. showed that the
interacts with pure DPPC bilayers, it is only with 50% that we inhibition constant of glutathione-mediated destruction of heme
observed significant changes in the ratio h+1/h0, a parameter that in ghost membranes by AQ is 25% greater than by chloroquine
increases to unity with increasing disordering. This indicates that (Famin et al., 1999). It was supposed that this difference might be
the bilayer surface charge has a crucial role in the AQ-bilayer related to different locations of AQ, chloroquine, and FPIX within
interaction. the membranes. It was proposed that chloroquine is located
mainly in the head group region of the lipid bilayer, but not
3.3. Possible role of drug-membrane interactions to the antimalarial intercalated into the phospholipids (Ginsburg and Demel, 1983;
mechanism of action Kursch et al., 1983 Zidovetzki et al., 1989). However, AQ is much
more hydrophobic than chloroquine (the partition coefficient of
The amphiphilic or hydrophobic character of many pharmaco- AQ in octanol is 359% greater than that of chloroquine, according to
logically active compounds confers them different types of Bray et al., 1996) and thus it may intercalate in the bilayer, as we
association, such as self-association and specific and/or non- showed in the present paper. In this way, the non-specific
specific binding to proteins and lipid membranes. Usually, the interaction between lipids and AQ, as other antimalarial drugs,
plasma membrane is one of their sites of action and, even if the may have an important role in their overall and complex
target is intracellular, the interaction with this first barrier plays an mechanism of action.
important role (Schreier et al., 2000). Since several antimalarials
are amphiphilic, such as AQ, chloroquine, primaquine, quinacrine, 4. Conclusions
mefloquine, and quinine, drug-membrane interactions may
contribute to their whole mechanism of action. Indeed, inter- We showed that AQ is able to interact non-specifically with
actions between halofantrine and lipid bilayers of PC and PE were DPPC/DPPS lipid bilayers affecting the gel phase more than the
observed by DSC measurements (Lim and Go, 1999). Halofantrine lipid fluid phase. AQ seems to influence more organized
caused a broadening of the phase transition of PC bilayers and a lipid domains, since the fluid phase of the model membrane
decrease in both enthalpy and transition temperature. A similar investigated was not affected. This AQ-bilayer interaction is driven
effect of halofantrine was observed with PE. Our results also by a combination of electrostatic forces and hydrophobic effects.
showed a decrease in the transition temperature caused by AQ. In Firstly, the approximation between the drug and the bilayer
another study, the effects caused by halofantrine on the phase surface is driven by electrostatic forces between the positively
transition of DPPC bilayers were compared with those caused charged amino group of AQ and the negatively charged PS
by quinine and mefloquine (Lim and Go, 1995). Similar to headgroups. Once close the bilayer, hydrophobic effects also
halofantrine, the latter drugs also interacted with DPPC bilayers, take place. Our results indicate that AQ binding to the binary
changing the phase transition and decreasing the transition DPPC/DPPS system shifts the equilibrium constant between
temperature. Pajeva et al. showed that quinacrine was also able the disordered DPPC-rich and the ordered DPPS-rich domains
R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78 77

towards the DPPC-rich one. We suggest that AQ is located Dyszy, F., Pinto, A.P.A., Araújo, A.P.U., Costa-Filho, A.J., 2013. Probing the interaction
completely inside the bilayer with the positively charged amino of brain fatty acid binding protein (B-FABP) with model membranes. PLoS One 8,
e60198.
group and the OH of the phenol group pointing outwards to the Earle, K.A., Moscicki, J.K., Ge, M., Budil, D.E., Freed, J.H., 1994. 250-GHz electron spin
polar headgroups of the bilayer. resonance studies of polarity gradients along the aliphatic chains in
phospholipid membranes. Biophys. J. 66, 1213–1221.
Elslager, E.F., Gold, E.H., Tendick, F.H., Werbel, L.M., Worth, D.F., 1964. Amodiaquine
Conflict of interest N-oxides and other 7-chloro-4-aminoquinoline N-oxides. J. Heterocycl. Chem. 1,
6–12.
The authors declare that there are no conflicts of interest. Escribá, P.V., 2006. Membrane-lipid therapy: a new approach in molecular
medicine. Trends Mol. Med. 12, 34–43.
Fajer, P., Watts, A., Marsh, D., 1992. Saturation transfer continuous wave saturation,
Transparency document and saturation recovery electron spin resonance studies of chain-spin labeled
phosphatidylcholines in the low temperature phases of dipalmitoyl
phosphatidylcholine. Biophys. J. 61, 879–891.
The Transparency document associated with this article can be
Famin, O., Krugliak, M., Ginsburg, H., 1999. Kinetics of inhibition of glutathione-
found in the online version. mediated degradation of ferriprotoporphyrin IX by antimalarial drugs.
Biochem. Pharmacol. 58, 59–68.
Acknowledgments Fitch, C.D., 2004. Ferriprotoporphyrin IX phospholipids, and the antimalarial actions
of quinoline drugs. Life Sci. 74, 1957–1972.
Freed, J.H., 1976. Theory of slow tumbling ESR spectra for nitroxides. In: Berliner, L.J.
This research was supported by USP, FAPESP (Grants No. (Ed.), Spin Labeling: Theory and applications. Academic Press, New York, pp.
2010/17662-8, 2011/1755-1 and 2012/20367-3) and CNPQ. The 53–132 (Chapter 3).
Ge, M., Freed, J.H., 1998. Polarity profiles in oriented and dispersed
authors thank Professor P. Ciacanglini and Professor M. T. Lamy for phosphatidylcholine bilayers are different: an electron spin resonance study.
allowing access to the DSC equipments. Biophys. J. 74, 910–917.
Ge, M., Freed, J.H., 2003. Hydration, structure, and molecular interactions in the
headgroup region of dioleoylphosphatidylcholine bilayers: an electron spin
Appendix A. Supplementary data resonance study. Biophys. J. 85 (6), 4023–4040.
Ginsburg, H., Demel, R.A., 1983. The effect of ferriprotoporphyrin IX and chloroquine
Supplementary data associated with this article can be on phospholipid monolayer and the possible implications to antimalarial
activity. Biochim. Biophys. Acta 732, 316–319.
found, in the online version, at http://dx.doi.org/10.1016/j.
Ginsburg, H., Famin, O., Zhang, J., Krugliak, M., 1998. Inhibition of glutathione-
chemphyslip.2014.12.003. dependent degradation of heme by chloroquine and amodiaquine as a
possible basis for their antimalarial mode of action. Biochem. Pharmacol. 56,
1305–1313.
References Gordon, V.D., Beales, P.A., Zhao, Z., Blake, C., Mackintosh, F.C., Olmsted, P.D., Cates, M.E.,
Egelhaaf, S.U., Poon, W.C.K., 2006. Lipid organization and the morphology of
C. Altenbach, 2014. LabVIEW programs for the analysis of EPR Data https://sites. solid-like domains in phase-separating binary lipid membranes. J. Phys. Condens.
google.com/site/altenbach/Home. Matter 18, L415–L420.
Ahmad, P., Mellors, A., 1978. Nuclear magnetic resonance studies in liposomes: effects Ghosh, A.K., Basu, R., Dey, S., Das, S., Nayak, N.P., Barat, B., Nandy, P., 1995. Lipid-
of steroids on lecithin fatty acyl chain mobility. J. Membr. Biol. 41, 235–247. disordering effect of aspirin on the liposomal membrane of dipalmitoyl
Alaouie, A.M., Smirnov, A.I., 2006. Ultra-stable temperature control in EPR phosphatidyl choline – a fluorescence anisotropy study. Colloid Surf. B 4, 309–311.
experiments: thermodynamics of gel-to-liquid phase transition in spin-labeled Gourley, D.H.R.,1971. Interaction of Drugs with Cells, Springfield.
phospholipid bilayers and bilayer perturbations by spin labels. J. Magn. Reson. Hamilton, K.S., Barber, K.R., Davis, J.H., Neil, K., Grant, C.W.M., 1991. Phase behaviour
182, 229–238. of amphotericin B multilamellar vesicles. Biochim. Biophys. Acta-Biomembr.
Bach, D., Wachtel, E., 2003. Phospholipid/cholesterol model membranes: formation 1062, 220–226.
of cholesterol crystallites. Biochim. Biophys. Acta 1610, 187–197. Heimburg, T., 2007. Thermal Biophysics of Membranes. Wiley-VHC, Weinheim.
Basso, L.G.M., Rodrigues, R.Z., Naal, R.M.Z.G., Costa-Filho, A.J., 2011. Effects of the Hsiao, L.L., Howard, R.J., Aikawa, M., Taraschi, T.F., 1991. Modification of host cell
antimalarial drug primaquine on the dynamic structure of lipid model membrane lipid composition by the intra-erythrocytic human malaria parasite
membranes. Biochim. Biophys. Acta 1808, 55–64. Plasmodium falciparum. Biochem. J. 274, 121–132.
Benatti, C.R., Epand, R.M., Lamy, M.T., 2007. Low cholesterol solubility in DODAB Huang, C.H., 1977. A structural model for the cholesterol–phosphatidylcholine
liposomes. Chem. Phys. Lipids 145, 27–36. complexes in bilayer membranes. Lipids 12, 348–356.
Biaggi, M.H., Riske, K.A., Lamy-Freund, M.T., 1997. Melanotropic peptides-lipid Jacob, H., Winterhalter, K., 1970. Unstable hemoglobins: the role of heme loss in
bilayer interaction: comparison of the hormone alpha-MSH to a biologically Heinz body formation. Proc. Natl. Acad. Sci. U. S. A. 65, 697–701.
more potent analog. Biophys. Chem. 67, 139–149. Jain, M.K., Wu, N.M., 1977. Effect of small molecules on the dipalmitoyl lecithin
Bittman, R., 2004. The 2003 ASBMB-Avanti Award in lipids address: applications of liposomal bilayer: III. Phase transition in lipid bilayer. J. Membr. Biol. 34,
novel synthetic lipids to biological problems. Chem. Phys. Lipids 129, 111–131. 157–201.
Blauer, G., Akkawi, M., Bauminger, E.R., 1993. Further evidence for the interaction of Könczöl, F., Farkas, N., Dergez, T., Belágyi, J., LÅrinczy, D., 2005. Effect of tetracaine on
the antimalarial drug amodiaquine with ferriprotoporphyrin IX. Biochem. model and erythrocyte membranes by DSC and EPR. J. Therm. Anal. Calorim. 82,
Pharmacol. 46, 1573–1576. 201–206.
Bray, P.G., Hawley, S.R., Mungthin, M., Ward, S.A., 1996. Physicochemical properties Kursch, B., Lüllmann, H., Mohr, K., 1983. Influence of various cationic amphiphilic
correlated with drug resistance and the reversal of drug resistance in drugs on the phase-transition temperature of phosphatidylcholine liposomes.
Plasmodium falciparum. Mol. Pharmacol. 50, 1559–1566. Biochem. Pharmacol. 32, 2589–2594.
Budai, M., Szabó, Z., Szogyi, M., Gróf, P., 2003. Molecular interactions between Lee, A.G., 2004. How lipids affect the activities of integral membrane proteins.
DPPC and morphine derivatives: a DSC and EPR study. Int. J. Pharm. 250, Biochim. Biophys. Acta 1666, 62–87.
239–250. Lim, L.Y., Go, M.L., 1995. A differential scanning calorimetry study of the interaction
Budil, D.E., Lee, S., Saxena, S., Freed, J.H., 1996. Nonlinear-least-squares analysis of of the antimalarial agent halofanthrine with dipalmitoyl phosphatidyl choline
slow-motion EPR spectra in one and two dimensions using a modified bilayers. Chem. Pharm. Bull. 43, 2226–2231.
Levenberg–Marquardt algorithm. J. Magn. Reson. Ser. A 120, 155–189. Lim, L.Y., Go, M.L., 1999. The antimalarial agent halofantrine perturbs
Chiang, Y.W., Shimoyama, Y., Feigenson, G.W., Freed, J.H., 2004. Dynamic molecular phosphatidylcholine and phosphatidylethanolamine bilayers: a differential
structure of DPPC–DLPC–cholesterol ternary lipid system by spin-label electron scanning calorimetric study. Chem. Pharm. Bull. 47, 732–737.
spin resonance. Biophys. J. 87 (4), 2483–2496. Luna, E.J., McConnell, H.M., 1977. Lateral phase separations in binary mixtures of
Cohen, P., 2002. Protein kinases – the major drug targets of the twenty-first phospholipids having different charges and different crystalline structures.
century? Nat. Rev. Drug Discov. 1, 309–315. Biochim. Biophys. Acta 470, 303–316.
Costa-Filho, A.J., Shimoyama, Y., Freed, J.H., 2003. A 2D-ELDOR study of the liquid Mainali, L., Hyde, J.S., Subczynski, W.K., 2013. Using spin-label W-band EPR to study
ordered phase in multilamellar vesicle membranes. Biophys. J. 84 (4), membrane fluidity profiles in samples of small volume. J. Magn. Reson. 226,
2619–2633. 35–44.
Couto, S.G., Cristina Nonato, M., Costa-Filho, A.J., 2011. Site directed spin labeling Mason, R.P., Rhodes, D.G., Herbette, L.G., 1991. Reevaluating equilibrium and kinetic
studies of – dihydroorotate dehydrogenase N-terminal extension. Biochem. binding parameters for lipophilic drugs based on a structural model for drug
Biophys. Res. Commun. 414, 487–492. interactions with biological membranes. J. Med. Chem. 34, 869–877.
Duarte, E.L., Oliveira, T.R., Alves, D.S., Micol, V., Lamy, M.T., 2008. On the interaction McMullen, T.P., Lewis, R.N., McElhaney, R.N., 2000. Differential scanning
of the anthraquinone barbaloin with negatively charged DMPG bilayers. calorimetric and Fourier transform infrared spectroscopic studies of the effects
Langmuir 24, 4041–4049. of cholesterol on the thermotropic phase behavior and organization of a
Duwe, H.P., Kaes, J., Sackmann, E., 1990. Bending elastic moduli of lipid bilayers: homologous series of linear saturated phosphatidylserine bilayer membranes.
modulation by solutes. J. Phys. 51, 945–961. Biophys. J. 79, 2056–2065.
78 R.P. Barroso et al. / Chemistry and Physics of Lipids 186 (2015) 68–78

Meirovitch, E., Nayeem, A., Freed, J.H., 1984. Analysis of protein-lipid interactions Sevcsik, E., Pabst, G., Richter, W., Danner, S., Amenitsch, H., Lohner, K., 2008.
based on model simulations of electron spin resonance spectra. J. Phys. Chem. Interaction of LL-37 with model membrane systems of different complexity:
88 (16), 3454–3465. influence of the lipid matrix. Biophys. J. 94, 4688–4699.
Naisbitt, D.J., Ruscoe, J.E., Williams, D., O'Neill, P.M., Pirmohamed, M., Park, B.K., 1997. Seydel, J., Wiese, M., 2002. Drug-membrane interactions: analysis, drug
Disposition of amodiaquine and related antimalarial agents in human distribution, modeling. Wiley-VHC, Weinheim.
neutrophils: implications for drug design. J. Pharmacol. Exp. Ther. 280, 884–893. Sherman, I.W., 1977. Amino acid metabolism and protein synthesis in malarial
O’Neill, P.M., Bray, P.G., Hawley, S.R., Ward, S.A., Park, B.K., 1998. 4-Aminoquinolines parasites. Bull. World Health Organ. 55, 265–276.
– past, present, and future: a chemical perspective. Pharmacol. Ther. 77, 29–58. Schultz, Z.D., Pazos, I.M., McNeil-Watson, F.K., Lewis, E.N., Levin, I.W., 2009.
Olliaro, P., Nevill, C., LeBras, J., Ringwald, P., Mussano, P., Garner, P., Brasseur, P., 1996. Magnesium-induced lipid bilayer microdomain reorganizations: implications
Systematic review of amodiaquine treatment in uncomplicated malaria. Lancet for membrane fusion. J. Phys. Chem. B 113, 9932–9941.
348, 1196–1201. Subczynski, W.K., Markowska, E., Gruszecki, W.I., Sielewiesiuk, J., 1992. Effects of
Overington, J.P., Al-Lazikani, B., Hopkins, A.L., 2006. How many drug targets are polar carotenoids on dimyristoylphosphatidylcholine membranes: a spin-label
there? Nat. Rev. Drug Discov. 5, 993–996. study. Biochim. Biophys. Acta 1105, 97–108.
Pajeva, I.K., Wiese, M., Cordes, H.P., Seydel, J.K., 1996. Membrane interactions of Swamy, M.J., Ciani, L., Ge, M., Smith, A.K., Holowka, D., Baird, B., Freed, J.H., 2006.
some catamphiphilic drugs and relation to their multidrug-resistance- Coexisting domains in the plasma membranes of live cells characterized by
reversing ability. J. Cancer Res. Clin. 122, 27–40. spin-label ESR spectroscopy. Biophys. J. 90 (12), 4452–4465.
Pentak, D., Sułkowski, W.W., Sułkowska, A., 2008. Calorimetric and EPR studies of Usman, M., Siddiq, M., 2013. Surface and micellar properties of chloroquine
the thermotropic phase behavior of phospholipid membranes. J. Therm. Anal. diphosphate and its interactions with surfactants and human serum albumin. J.
Calorim. 93, 471–477. Chem. Therm. 58, 359–366.
Pussard, E., Verdier, F., 1994. Antimalarial 4-aminoquinolines: mode of action and Van Dijck, P.W., 1979. Negatively charged phospholipids and their position in the
pharmacokinetics. Fundam. Clin. Pharmacol. 8, 1–17. cholesterol affinity sequence. Biochim. Biophys. Acta 555, 89–101.
Sá, J.M., Twu, O., Hayton, K., Reyes, S., Fay, M.P., Ringwald, P., Wellems, T.E., 2009. WHO, 2010. Global report on antimalarial drug efficacy and drug resistance:
Geographic patterns of Plasmodium falciparum drug resistance distinguished by 2000–2010, Drug Resistance Updates.
differential responses to amodiasackquine and chloroquine. Proc. Natl. Acad. WHO, 2011. World Malaria Report 2011.
Sci. U. S. A. 106, 18883–18889. Yokoyama, H., Ikeda, K., Wakabayashi, M., Ishihama, Y., Nakano, M., 2013. Effects of
Sackmann, E., 2006. Thermo-elasticity and adhesion as regulators of cell membrane lipid membrane curvature on lipid packing state evaluated by isothermal
architecture and function. J. Phys. Condens. Matter 18, R785–R825. titration calorimetry. Langmuir 29, 857–860.
Santos, M.S., Del Lama, M.P.F.M., Ito, A.S., Naal, R.M.Z.G., 2014. Binding of chloroquine to Zhao, L., Feng, S.-S., Kocherginsky, N., Kostetski, I., 2007. DSC and EPR investigations
ionic micelles: effect of pH and micellar surface charge. J. Lumin. 147, 49–58. on effects of cholesterol component on molecular interactions between
Schneider, D.J., Freed, J.H., 1989. Calculating slow motional magnetic resonance paclitaxel and phospholipid within lipid bilayer membrane. Int. J. Pharm. 338,
spectra: a user’s guide. In: Berliner, L.J., Reuben, J. (Eds.), Biological Magnetic 258–266.
Resonance, vol. 8. Plenum Publisher Corporation, New York, pp. 1–76. Zidovetzki, R., Sherman, I.W., Atiya a De Boeck, H., 1989. A nuclear magnetic
Schreier, S., Malheiros, S., de Paula, V., 2000. Surface active drugs: self-association resonance study of the interactions of the antimalarials chloroquine quinacrine,
and interaction with membranes and surfactants: physicochemical and quinine and mefloquine with dipalmitoylphosphatidylcholine bilayers. Mol.
biological aspects. Biochim. Biophys. Acta 1508, 210–234. Biochem. Parasitol. 35, 199–207.

You might also like