You are on page 1of 10

Journal of Molecular Graphics and Modelling 111 (2022) 108110

Contents lists available at ScienceDirect

Journal of Molecular Graphics and Modelling


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

The effect of external salts on the aggregation of the multiheaded


surfactants: All-atom molecular dynamics studies
Sandeep Dash 1, Unmesh D. Chowdhury 1, B.L. Bhargava *
School of Chemical Sciences, National Institute of Science Education & Research-Bhubaneswar, HBNI, P.O.Jatni, Khurda, Odisha, 752 050, India

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

Keywords: Tailoring the molecular design of the surfactants leads to changes in the aggregation properties. The role of
Molecular dynamics external salts on the aggregation properties of the multiheaded surfactants is investigated using molecular dy­
Micelles namics simulations. The multiheaded surfactants show differential aggregation properties on addition of external
Multiheaded surfactants
salts, as reported earlier from experimental studies. We have modelled the multiheaded surfactants to study the
Micellization
effect of external salts (potassium bromide and sodium salicylate) at three different concentrations using the all-
atom modelling and explicit solvation. The influence of external salts on the hydration and aggregation pro­
pensity, hydrogen bonding, and the structural characteristics of the surfactant aggregates are probed using
various analyses across the four groups of multiheaded surfactants. The larger salicylate ion masks the repulsion
between the cationic head groups and acts as an effective promoter of aggregation.

1. Introduction aggregation properties differ through the changing of the number of


head groups [15].
Surfactants are a class of amphiphilic molecules having varied ap­ One of the well known examples of the cationic surfactants, (sur­
plications in the field of biology and chemistry [1]. The “amphiphilicity” factants with a positively charged head group) is the cetyl­
refers to the presence of both hydrophobic and hydrophilic groups in the trimethylammonium bromide, which is commonly known as CTAB [16,
surfactant molecule [2]. Surfactants form a part of our daily lives being 17]. This is also an example of an ionic micelle, which forms aggregates
the primary components of the soaps and detergents that we use. They of (CTA+) that are stabilized by bromide (Br− ) counterions. Aqueous
are also present in our bodies, such as, in the Bile Salts, which are ex­ solutions of CTAB and similar surfactants have been extensively studied
amples of bio-surfactants that help in the digestion process [3]. Sur­ [18–21]. The usual behavior of these surfactants has been observed to
factants also act as antiseptics aiding the recovery of wounds faster [4, change with addition of external salts to their solutions. There are many
5]. They also play a role in the process of generation of medicinal experimental reports of the effect of sodium salicylate (NaSal) salt and
aerosols for respiration from nebulizers [6]. Surfactant based process of salicylic acid on the CTAB surfactant solutions [22–26], which show the
remediation of soil and water contaminated by organic contaminants increase in micellar growth and changes in the viscoelastic properties of
has proved to be faster and more effective than other technologies [7]. the system. Increase in the micelle formation in presence of alkali halide
The aggregation properties of the surfactant molecules differs through salts like KBr in aqueous CTAB solutions using neutron and light scat­
their changing architecture. Various kinds of surfactants have been tering experiments has also been reported [27].
discovered, which differs in their structures, charges and also on the Multiheaded variants of CTAB, having more than two head groups
overall architecture of the molecules [8–12]. A few examples of sur­ have been introduced by Haldar et al. [15,28] The multiple hydrophilic
factants having varying architecture are the gemini surfactants head groups are connected to a single hydrophobic tail in these variants.
(composed of two hydrophilic head groups and two hydrophobic tails With the increase in the number of head groups, the solubility of these
linked by a spacer at the head groups), bola amphiphiles (composed of surfactants in water is observed to increase with consequent decrease in
hydrophilic group at the both ends separated by a hydrophobic chain), aggregation. This is due to the increase in the repulsion between the
phospholipids etc [13,14]. Hence, for the multiheaded surfactants, the surfactants with increase in the charge on the cationic head groups.

* Corresponding author.
E-mail address: bhargava@niser.ac.in (B.L. Bhargava).
1
Equal contribution.

https://doi.org/10.1016/j.jmgm.2021.108110
Received 31 August 2021; Received in revised form 13 December 2021; Accepted 13 December 2021
Available online 17 December 2021
1093-3263/© 2021 Elsevier Inc. All rights reserved.
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

Small angle neutron scattering (SANS) studies [15] have also confirmed
that the addition of external salts to the aqueous solutions of these
surfactants increases the aggregation of these surfactants, by decreasing
the repulsions between the similarly charged head-groups. In the recent
years, molecular dynamics and other computational methods have been
extensively used to study different micellar systems showing versatile
properties [29–31]. Computational studies of the surfactant aggregation
have also been carried out [32–42]. Coarse grained simulations of
multiheaded variants of CTAB and cetylpyridinium bromide (CPB)
surfactant systems have shown that the aggregation number in the so­
lutions decrease with increase in the number of head groups which has
been further confirmed from all atoms molecular dynamics simulations
[43,44]. Our study is focused on the effect of the external salts on
micellar aggregation. When salts are added to the surfactant solution,
the repulsion between the hydrophobic headgroups is decreased which
results in variation of the shape and size of the aggregates. The effect of
the added salts in the aggregation of the gemini surfactants is reported
earlier [45,46]. Micelles may change from global to rodlike or wormlike
with the addition of inorganic and organic salts such as the salicylate
and the tosylate [25,47–49]. The addition of salt in ionic surfactant
solutions changes the aggregation behavior which is vital for many in­
dustrial processes including emulsification. Thus, the understanding of
the molecular level interactions involved in the surfactant aggregation Fig. 1. Structure of multiheaded surfactants used in our study.
with the addition of external salts is important [50].
NaSal is used to describe sodium salicylate salt.
2. Methodology and simulation details
3. Results and discussions
The all-atom molecular dynamics simulations of aqueous solutions of
multiheaded surfactants (with one to four headgroups) were performed 3.1. Radial distribution function (RDF)
using GROMACS (version 5.1.4) software [51]. The simulations were
carried out at various concentrations of two different salts, potassium The radial distribution function, g(r), determines the local in­
bromide (KBr) and sodium salicylate (NaSal) added to the solution, homogeneity arising due to the intermolecular interactions. In the
accounting for a total of 28 different systems. following discussion, the cation is represented by the positively charged
The OPLS-AA [52] force field has been used to model the surfactant nitrogen atom of the pyridinium ring of the headgroup and the water
molecules along with the simple point charge extended (SPCE) [53] molecules are represented by their oxygen atoms.
water model. All the bonds involving hydrogen atoms were constrained The added salts aid in decreasing the repulsion between the head
using the LINCS algorithm [54]. The Nosé-Hoover thermostat [55] with groups, thus allowing the cations to come closer. This is evident in the
1.0 ps time constant and 298 K reference temperature, and case of the 1H system shown in Fig. 2, where the peak amplitude cor­
Parrinello-Rahman barostat [56] with 0.5 ps time constant and 1 bar responding to the first maximum increases with increase in the salt
reference pressure were used to control the temperature and pressure concentration. The addition of the external Sal− ions increases the
respectively. The van der Waals and Coulomb interaction were spatial organization of the cation head groups around themselves. The
computed up to a cutoff distance of 1.3 nm. The long ranged interactions external KBr salts are less effective in reducing the cation – cation re­
were modelled using the Particle Mesh Ewald (PME) method [57]. pulsions and bringing the cations together.
The systems were initially energy minimized using the steepest The radial distribution of cations around themselves in the other
descent method and then equilibrated in canonical ensemble (constant systems are shown in supplementary information (Fig. S1 – S3). It is
NVT) for 4 ns followed by another equilibration of 4 ns in isothermal- observed that the amplitude of the first maxima decreases with increase
isobaric (constant NPT) ensemble. The production runs were carried in the number of head groups. This can be attributed to the fact that the
out in constant NPT ensemble for 100 ns and the last 20 ns trajectory was increase in the number of head groups increases the electrostatic
used for the subsequent analyses. A time step of 1 fs was used for all the repulsion between them thus decreasing the probability of finding other
simulations, and the position and velocities were stored every 0.5 ps. cations around them. The first maxima is observed between 0.7 and 0.8
Visual Molecular Dynamics package (VMD) [58] and PyMol [59] nm in all the systems, which may be considered as the distance between
were used to render the images. The analyses were carried out using the a cation and the first shell of other cations around it. Addition of NaSal
gmx tools and in-house codes using the MDAnalysis Python package increases the organization of cations around themselves in all the sys­
[60]. The schematic of the multiheaded surfactants are provided in tems including the 4H solution, unlike with the addition of KBr. This
Fig. 1. observation and the greater amplitude of the first maxima in the RDFs in
case of NaSal addition hints about better efficiency of NaSal in the
2.1. System description promotion of aggregation compared to KBr.
The radial distribution of the bromide ions around the cations in 1H
A total of 28 systems were prepared for the study. The number of salt system are shown in Fig. 3.
pairs added to the systems has been maintained at a constant fraction of In solutions with KBr salt, the intensity of the first maxima, present at
the cation molecules across the multiheaded systems for proper com­ around 0.3 nm, decreases with increase in the number of head groups as
parison of the properties. The details of the studied systems are shown in given in the supplementary information (Fig. S4 – S6). The second
Table 1. The external salts are added using the gmx code. The ratio of maxima are observed at around 0.5 nm and the intensity of these peaks
cations to water is maintained at 1:100 in all the systems. We have used also decreases with increasing number of head groups. The interaction
the shorthand notation ‘nH’ to describe various multiheaded surfactant between the cations and anions does not vary significantly with varying
systems, where ‘n’ refers to the number of head groups in the cation.

2
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

Table 1
Details of the aqueous solutions of the multiheaded surfactants used in the study.
Number of Head No. of No. of No. of Final Box No. of
Groups and salt surfactant waters added Length Atoms
concentration ion pairs salt (nm)
pairs

1H 100 10 000 — 7.136 36 400


1H (0.05 mol 100 10 000 KBr (5) 7.139 36 710
ratio)
1H (0.25 mol 100 10 000 KBr (25) 7.145 36 750
ratio)
1H (0.50 mol 100 10 000 KBr (50) 7.159 36 800
ratio)
1H (0.05 mol 100 10 000 NaSal 7.152 36 780
ratio) (5)
1H (0.25 mol 100 10 000 NaSal 7.167 37 100
ratio) (25)
1H (0.50 mol 100 10 000 NaSal 7.196 37 500
ratio) (50)
2H 100 10 000 — 7.285 38 700
2H (0.05 mol 100 10 000 KBr (5) 7.281 38 710
ratio)
2H (0.25 mol 100 10 000 KBr (25) 7.292 38 750
ratio)
2H (0.50 mol 100 10 000 KBr (50) 7.307 38 800
ratio) Fig. 2. The RDF of the cations around themselves in 1H systems in the presence
2H (0.05 mol 100 10 000 NaSal 7.287 38 780 of added salt.
ratio) (5)
2H (0.25 mol 100 10 000 NaSal 7.314 39 100
ratio) (25)
2H (0.50 mol 100 10 000 NaSal 7.339 39 500
ratio) (50)
3H 100 10 000 — 7.411 40 700
3H (0.05 mol 100 10 000 KBr (5) 7.422 40 710
ratio)
3H (0.25 mol 100 10 000 KBr (25) 7.440 40 750
ratio)
3H (0.50 mol 100 10 000 KBr (50) 7.439 40 800
ratio)
3H (0.05 mol 100 10 000 NaSal 7.424 40 780
ratio) (5)
3H (0.25 mol 100 10 000 NaSal 7.443 41 100
ratio) (25)
3H (0.50 mol 100 10 000 NaSal 7.476 41 500
ratio) (50)
4H 60 6000 — 6.402 25 980
4H (0.05 mol 60 6000 KBr (5) 6.408 25 986
ratio)
4H (0.25 mol 60 6000 KBr (25) 6.409 26 010
ratio)
4H (0.50 mol 60 6000 KBr (50) 6.420 26 040
ratio)
4H (0.05 mol 60 6000 NaSal 6.402 26 028
ratio) (5)
4H (0.25 mol 60 6000 NaSal 6.416 26 220
ratio) (25) Fig. 3. The RDFs of the anions around the cation in 1H systems in the presence
4H (0.50 mol 60 6000 NaSal 6.441 26 460 of added salt.
ratio) (50)

cation. In case of multiple head groups, it is observed that the amplitude


KBr salt concentration in systems with more head groups. The variation of first and second peaks are comparable in case of NaSal systems unlike
is more pronounced in case of 1H systems. the KBr systems which exhibit a shorter second peak. Significant pop­
In the presence of NaSal salt, the cation – anion RDFs exhibit ulation of Sal− ions around the cations reduce the repulsion between the
different behavior. The cation – anion RDF refers to the distribution of head groups aiding in the aggregation of cations.
bromide anions around cation head groups. It can be noticed that there The RDFs of the water molecules around the cations show a gradual
is a significant increase in the amplitude of the first maxima with the increase with the increase in the number of head groups (Fig. S7 – S10,
addition of NaSal salt, compared to the no salt case in 1H systems. The Supplementary Information). The differences in the RDFs in case of
Sal− ions are predominantly present around the cations at around 0.5 systems with KBr salts are not very prominent, unlike those with NaSal
nm distance. The amplitude of the second maxima present at around 1.2 salt. The presence of the polyatomic aryl group in the salicylic anion
nm are small compared to the first ones in the 1H systems. The RDFs in reduces the interaction between the water molecules and the cationic
the presence of NaSal are qualitatively similar to those with KBr, but the head groups leading to the decrease in the value of the corresponding
positioning of the first and second coordination shells of anions are cation – water RDFs with the increase in the concentration of NaSal salt.
different. In the cation – anion RDFs of the solutions with 2H, 3H and 4H The difference in the organization of the water molecules around the
given in the supplementary information, one can observe the multiple anions (Br− and Sal− ) is evident from the anion – water RDFs corre­
peaks that are arising due to the presence of multiple head groups on the sponding to the 1H systems shown in Fig. 4. Fig. S11 – S13,

3
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

Fig. 4. The radial distribution of the water molecules around the anions in the
1H systems in the presence of added salt. Fig. 5. The SDF of (a) Br− ions around the cation head group in 1H system with
KBr (50) and (b) Sal− ions around the head group in 1H system with NaSal (50).
corresponding to 2H, 3H and 4H systems are given in the supplementary
information. In case of the KBr salt, all bromide ions in the solution are ● CH⋯O H-bond, where the pyridine ring C-atoms in the surfactant
considered. head groups act as the donors and the carboxylate O-atoms of Sal−
In case of solutions with KBr salt, there is a sharp first maxima at 0.3 ions are chosen as acceptors.
nm followed by comparatively broader second and third maxima at 0.5 ● OH⋯N H-bond, where the hydroxyl oxygen − atoms of Sal− ions act
nm and 0.7 nm respectively, pointing to well defined multiple solvation as donors and the pyridine N-atoms in the surfactant head groups are
shells of water around the anions. The peak intensities are found to chosen as acceptors.
gradually increase with the number of head groups. Water molecules are
more organized around anions compared to the cations. The peak in­ The hydrogen bonds are classified as strong H-bonds if the distance
tensity for the anions with the water in the Sal− systems are markedly between the donor and the acceptor is less than 0.22 nm and the angle
reduced. between the donor, hydrogen atom of the donor and the acceptor is
between 140◦ and 180◦ . The corresponding distance and angle values
3.2. Spatial distribution functions are 0.30 nm and 90◦ –180◦ , respectively for the weak H-bonds [61].
The average number of the hydrogen bonds observed in the surfac­
To elucidate the interactions between the anions and the cations tant - salicylate systems are given in Table 2. The higher concentration of
further, we have computed the spatial distribution functions (SDFs) of NaSal induces higher number of CH⋯O H-bonds in the solution as seen
the external ions (both Br− and Sal− ions) around the head groups of the in the Table 2. A snapshot of the arrangement of the surfactants and the
cations in 1H systems which are shown in Fig. 5. Sal− ions is given in Fig. 6(a) along with the pertaining atom types of the
The relative density of the anions is 2 times the average density. The head groups (Fig. 6(b)).
SDFs succinctly show the difference in the organization of the added ions It can be observed that there is one Sal− anion present in between the
around the surfactant head groups. The Sal− ions can reach the prox­ two head groups. This arrangement is similar to that of a salt bridge,
imity of the surfactant head groups better than the added KBr salts can which is one of the major structural motifs that provide stability to the
reach in their solutions. The solvation shell of the Br− ions around the protein structures [62–64]. This is further illustrated by the combined
head group of the cation is shifted away from the cation in the solution distance-angle distribution functions (CDFs) as shown in the case of 1H
with KBr salt compared to that in the NaSal case. Thus, it can be said that and 4H systems in Fig. 6(c) and (d) respectively. In the CDFs, the
the NaSal ions shield the cationic head groups better than the KBr, due to
the higher surface area of the salicylate ions.
Table 2
The average number of Hydrogen bonds per surfactant in solutions having NaSal
3.3. Hydrogen bonding
ions.

The Hydrogen bonding network plays a pivotal role in stabilizing the System No. of external salt Average number of CH⋯O weak hydrogen bond
micellar aggregates. The difference in the salting-in properties of the two 1H 5 0.19
salts can be attributed to the differences in their structures; KBr having 1H 25 0.97
1H 50 1.77
monatomic bromide ion and NaSal having polyatomic ion. From the
2H 5 0.19
RDF plot of the anion – water, it is observed that the higher surface area 2H 25 1.01
of the salicylic acid masks certain interactions which are otherwise 2H 50 2.08
present in case of the solutions with KBr salt. The following hydrogen 3H 5 0.17
bonding interactions were investigated: 3H 25 0.95
3H 50 1.69
4H 3 0.15
4H 15 0.69
4H 30 1.30

4
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

Fig. 7. RDFs of the salicylate oxygen atoms (OB, OC and OD) around the
surfactant cations.

Fig. 6. (a) The interaction of salicylate anion with the surfactant head group
showing the atom types of the carbon atoms. (b) The atom types of oxygen
atoms of the salicylate ion. OC and OB are the carboxylic oxygen atoms and the
OD is the hydroxyl oxygen atom. The combined distribution function (CDF) of
the distance and the angle (between the carboxylic group of the salicylate and
the ring C-atoms of the surfactant for the CH⋯O hydrogen bond) in case of (c)
1H system with 50 NaSal and (d) 4H system with 30 NaSal.

distance between the carboxylic oxygen atom and the hydrogen atoms
attached to the carbon atoms on the head group is considered and the
angle corresponds to the one made by C-atom of the headgroup, the
hydrogen attached to it and the carboxylic oxygen atom of the salicylate
ion. The CDFs show the formation of H-bonds between the head group
and the salicylate ions.
In order to understand the nature of the interaction of the surfactants
with the salicylate, we computed the site-site RDF of the oxygen atoms
of the carboxylate group and the hydroxyl group of salicylate with the
surfactant molecules, which are shown in Fig. 7.
From the figure, we can observe the differential distribution of the
salicylate oxygen atoms near the head groups of the cation. The
carboxylate oxygen atoms are more likely to be present around the
cations at a distance of around 0.25–0.3 nm. It is also evident that the
population of hydroxyl O-atoms around cation head groups is compar­
atively less. The difference in the population of O-atoms of the carbox­
ylate and hydroxyl group surrounding the cation decreases with
increasing number of head groups. This is because, the increase in the
number of cationic head groups reduces the free volume for the salicy­
late ions to be present near them despite the constancy in the number of Fig. 8. RDFs of the surfactant carbon atoms (CA, CE and CF) and salicylate
anions. anion carboxylate oxygen atoms (OB and OC).
The RDFs between the carbon atoms of the surfactant head group and
carboxyl oxygen atoms of the salicylate anion are shown in Fig. 8. per cation increases.
CA and CE are the carbon atoms adjacent to the nitrogen in the These results indicate that the dominant interactions between the
pyridine ring and CF is the carbon atom outside the pyridine ring that is cations and the Sal− ions are electrostatic in nature, similar to that of a
attached to the nitrogen (Fig. 6(a)). The important atom types of the salt bridge. We computed the distance and angle (N–C–O, where N is the
salicylate ion are shown in Fig. 6(b). The amplitudes of the first peaks cation nitrogen atom, C is the adjacent carbon atoms of the surfactant
are found to decrease with increase in the number of head groups. This is molecules and O is the carboxyl oxygen atoms of Sal− ions) distribution
because, the number of anions is constant in every system and hence the between the cationic head groups and the carboxylic oxygen in the
ratio of anions to head groups decreases as the number of head groups

5
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

salicylate ion to confirm the electrostatic salt bridge interaction. The


distributions are shown in Fig. 9(a).
The distance distribution peaks between 0.3 and 0.4 nm. The dis­
tribution of angle between the N–C–O atoms for the pairs having the N-
atom within 0.3 nm of the O-atom is shown in Fig. 9(b). Majority of the
pairs are found to be aligned such that the angles are between 50◦ –80◦
and the most probable angle is 69◦ . 3H and 4H systems are observed to
have nearly similar counts, whereas the counts increase for the 1H and
2H case as seen in Fig. 9(a)–(b). Various studies [62–64] have shown
that the formation of a salt bridge requires both the oppositely charged
residues to be within 0.4 nm from each other. We have also observed
similar distances between the positively charged surfactant and the
negatively charged Sal− ions. Therefore, the interaction between the
cations and Sal− ions are predominantly electrostatic in nature and are
similar to the salt bridges observed in proteins. These interactions are
observed only in case of Sal− ions due to the presence of carboxylate
group but are not found in case of KBr salt.

3.4. Hydration index

Hydration Index is a measure of the propensity of alkyl chain being


exposed to water or being ‘hydrated’. It is defined as the ratio of the
number of water molecules surrounding the surfactant in an aggregate
to the those around a single surfactant molecule in the bulk.
CNaggregate
Hydration Index = (1)
CNbulk
The hydration index for the 1H systems at various salt concentrations
are shown in Fig. 10A-(a). The hydration indices for the 2H, 3H and 4H
surfactants are shown in Figures (S14 – S16) in the supplementary
information.
In the figure, the first carbon is the one connected to the head group.
This carbon atom is exposed to the solvent and thus has higher value of
hydration index. We can see that the hydration index decreases across

Fig. 10. A-(a) The hydration index for the surfactants in the presence of KBr
and NaSal salts in case of the 1H headgroups. A-(b) The hydration index for the
surfactants at highest concentration of KBr and the NaSal. B and C are the
snapshots of the aggregates formed in 3H and 4H systems with the cation head
groups shown in cyan and tail groups shown in blue. The structure of the ag­
gregates exhibit the solvent exposure of the cation head group. (For interpre­
tation of the references to colour in this figure legend, the reader is referred to
the Web version of this article.)

the alkyl chain up to the 13th carbon atom and then start increasing for
the three carbon atoms towards the terminus, suggesting the exposure of
the last three carbon atoms to the solvent. This shows that the aggregates
formed in the solutions are not perfectly spherical with all the tail groups
buried inside.
Hydration index is observed to increase with an increase in the
number of head groups in all of the systems, which indicates greater
exposure to water molecules in case of systems with more head groups
resulting in the decreased cluster formation in 3H and 4H systems [43].
From Fig. 10A-(a), it can be observed that the hydration indices decrease
with increase in the salt concentration. The variation of hydration
indices with concentration is significant in case of NaSal systems
whereas the differences are less pronounced in case of KBr systems. The
Fig. 9. (a) The distance distribution of the pyridine N-atoms and the carboxylic
O-atom and (b) the angle distribution of the N–C–O angle. significant decrease observed in NaSal systems may be attributed to the

6
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

masking of carbon chain from the water molecules by the polyatomic


salicylate ions.
The variation of hydration index with head groups in the presence of
the external salt is compared in Fig. 10A-(b). The hydration indices are
observed to be lower in systems with NaSal compared to those with KBr
salt, which further confirms the increased effect of NaSal salt on the
hydration of cations.

3.5. Structural characteristics

3.5.1. Aggregation number


Through the course of the simulation, the monomers present in the
system come together and form aggregates. The aggregates are formed
in such a way that the alkyl chains are present inside the aggregates and
the head groups are facing outside, interacting with the solvent as shown
in Fig. 11.
To quantify the aggregates, we have determined the aggregation
numbers, i.e. the number of aggregates present in one cluster. This was
done by counting the neighbouring molecules present within a certain
cut-off range, say 1 nm, and then considering all the neighbours as one Fig. 12. The distribution of the aggregation numbers for the 1H surfactants in
single cluster. A minimum of 10 surfactants were considered to be the the presence of the external salts.
threshold to be considered as a cluster.
Fig. 12 shows the aggregation numbers for the 1H surfactant systems for the system with no added salt. The aggregation numbers are also
in the presence of external salts, KBr and NaSal. From the figure, it is dependent on the concentration of the external salts (both KBr and
evident that the aggregation is favored in the presence of NaSal salts and NaSal) which is in agreement with the earlier reports based on SANS
the concentration of salt correlates with the size of the aggregates. The experiments [15,28].
aggregation numbers observed for the NaSal salts are greater than that

Fig. 11. Snapshots of the representative aggregates


formed in multiheaded surfactant solutions in the
presence of salt. (a) and (b) are the snapshots of 1H
aggregates in KBr 50 and NaSal 50 respectively and,
(b) and (d) are the representative aggregates in case
of the 2H in KBr 50 and NaSal 50. Polar and
nonpolar regions of the cations are shown in cyan
and blue respectively. The water molecules and
hydrogens are omitted for clarity. (For interpreta­
tion of the references to colour in this figure legend,
the reader is referred to the Web version of this
article.)

7
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

The monotonic increase in the aggregation number with the con­ The aggregates of multiheaded surfactants are shown to exhibit changes
centration is not seen in case of cations with multiple head groups. in the semimajor and semiminor axis with the addition of external salts,
Nevertheless, the aggregation numbers are greater in systems with using SANS experiments [15]. The eccentricities which characterize the
added salt compared to those without. With increase in the number of shape of the aggregates have been obtained from the all-atom MD
head groups, one can observe the decrease in aggregation numbers, simulations.
which is in accordance with earlier reports [43]. In case of 2H system Fig. 13 shows the distribution of eccentricity values of aggregates in
with 50 KBr, the aggregation numbers observed are less than that for the various systems in the presence of the external salt. The KBr 25 systems
systems with 5 and 25 KBr (Fig. S19, supplementary information). are found to show more preference towards the ellipsoidal shape
However, the general trend of increase in aggregation numbers with salt whereas in the case of the NaSal systems, the Sal 50 shows the maximum
concentration is observed with NaSal in 2H systems (Fig. S20, supple­ preference towards the ellipsoidal shape. Thus, the surfactant aggre­
mentary information). The 3H and 4H systems show lower aggregation gates are having more penchant towards the ellipsoidal shape with the
numbers due to the increased electrostatic repulsions with increase in addition of the external salts in accordance with the observations from
the number of head groups. Addition of KBr results in slight increase in SANS experiments [15]. The ellipsoidal shape gives more room for the
the aggregation and the increase in salt concentration does not always head groups to negate the electrostatic interactions compared to the
promote more aggregation. NaSal on the other hand has the largest spherical shape.
aggregation number observed with highest concentration with very few We have also characterised the mean end-to-end distance of the
exceptions (Fig. S21 – S24, supplementary information). Thus, addition cations in the aggregates and the radius of gyration of the aggregates in
of NaSal facilitates the formation of larger sized aggregates in the the presence of salts. The mean end-to-end distance 〈Re− e〉 and the radius
solution. of gyration 〈Rg〉 are given by the following equations.
〈Re− e 〉 = 〈(rN − r1 )〉 (3)
3.5.2. End-to-end radius and eccentricity
The shape of the aggregate plays a critical role in the hydrodynamic ( )
behavior of the solution. The formation of the micelle results in a pro­ 1 ∑N
〈Rg 〉 = (rj − rc.m )2 0.5
(4)
found change in the radius of gyration and the end-to-end radius. The N j=1

mean end-to-end radius of the surfactants in the presence of various salts


are given in Table 3. For qualitative assessment of the shape of the ag­ 1 ∑N
rc.m = rj (5)
gregates, we use the eccentricity (ϵ), which is given by N j=1

Imin where r1 and rN denote the position vectors of the first and the last atoms
ϵ=1− (2)
Iavg of the chain backbone of the surfactant in the aggregate and rc.m is the
position vector of the center of mass of the chain. The structural char­
where Imin is the minimum of the moment of inertia along the x, y or z acteristics of the surfactants are given in Table 3.
axis and Iavg represents the average over the moment of inertia along the In the table, the two values present against the aggregation number
three axis. Eccentricity measures the deviation of a 3D ellipsoid from the column represent the aggregation numbers that have relative higher
true sphere. For a perfect sphere ϵ is 0 whereas the value of ϵ → 1 cor­ probability. The average end-to-end distance is found to decrease with
responds to a shape defying the sphere and going to an ellipsoidal shape. increase in the number of head groups on the cation which supports the

Table 3
Structural characteristics of the aggregates formed in aqueous solutions of
multiheaded surfactants in the presence of external salt.
System External 〈Re− e〉 〈Rg〉 〈ϵ〉 Aggregation
Salt [nm] [nm] number

1H no salt 2.29 0.79 0.31 20, 78


1H KBr(5) 2.28 0.78 0.33 78
1H KBr(25) 2.28 0.78 0.50 91
1H KBr(50) 2.29 0.79 0.30 28
1H Sal(5) 2.29 0.79 0.18 84
1H Sal(25) 2.26 0.78 0.23 92
1H Sal(50) 2.25 0.78 0.52 100
2H no salt 2.12 0.74 0.17 22
2H KBr(5) 2.05 0.73 0.06 44
2H KBr(25) 2.11 0.74 0.04 11, 18
2H KBr(50) 2.04 0.73 0.18 15, 67
2H Sal(5) 2.09 0.73 0.14 62
2H Sal(25) 2.08 0.72 0.11 18, 52
2H Sal(50) 2.07 0.73 0.18 59, 39
3H no salt 1.93 0.70 0.04 12,18
3H KBr(5) 1.92 0.69 0.07 31
3H KBr(25) 1.93 0.70 0.09 18, 24
3H KBr(50) 1.98 0.70 0.11 19, 17
3H Sal(5) 1.91 0.70 0.08 15, 18
3H Sal(25) 1.94 0.70 0.07 12, 22
3H Sal(50) 1.91 0.69 0.08 11, 29
4H no salt 1.76 0.69 0.04 11
4H KBr(3) 1.82 0.70 0.09 10
4H KBr(15) 1.82 0.67 0.12 15
4H KBr(30) 1.82 0.67 0.14 10
4H Sal(3) 1.81 0.69 0.07 14
4H Sal(15) 1.82 0.70 0.07 12
Fig. 13. The distribution of the eccentricity values of the aggregates in the 1H
4H Sal(30) 1.79 0.70 0.09 15, 20
system in the presence of (a) KBr and (b) NaSal salts.

8
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

observed decrease in their aggregation numbers. The mean radius of org/10.1016/j.jmgm.2021.108110.


gyration 〈Rg〉 is nearly similar in 1H and 2H systems. However, consis­
tent decrease in the radius of gyration is observed in systems with 3 and Abbreviation
4 headgroups, which is consistent with the earlier observations [28, 15,
44]. Presence of salt does not cause any regular variation in the end-to- CTAB Cetyltrimethylammonium bromide
end distance. Increase in salt concentration increases the aggregation NaSal Sodium Salicylate
but the highest concentration of added salt does not always increase the SANS small angle neutron scattering
aggregation number. A concentration of 25–50 number of external KBr RDF radial distribution function
and NaSal salts are found to be most effective to form the clusters. The SDF spatial distribution function
addition of salt is also found to increase the eccentricity of the aggre­ H-bond hydrogen bond
gates in general, which also correlates with the observation of aggre­
gation numbers. The experimental SANS study [15] reports that the
References
inclusion of the external salts leads to the change in the shape of the
aggregate with the least effect being found in the 2H case. It can be [1] D. Lombardo, M.A. Kiselev, S. Magazù, P. Calandra, Amphiphiles self-assembly:
observed from the table that the eccentricity is increasing in general basic concepts and future perspectives of supramolecular approaches, Adv.
with the increase in salt concentration, signifying the change in the Condens. Matter Phys. 2015 (2015).
[2] Y. Cao, S. Yang, Y. Li, J. Shi, Cooperative Organizations of Small Molecular
shape of the aggregates. Surfactants and Amphiphilic Block Copolymers: Roles of Surfactants in the
Formation of Binary Co-assemblies, 2021, p. e49. Aggregate.
4. Conclusions [3] P. Wilde, B. Chu, Interfacial & colloidal aspects of lipid digestion, Adv. Colloid
Interface Sci. 165 (1) (2011) 14–22.
[4] L. Caillier, E.T. de Givenchy, R. Levy, Y. Vandenberghe, S. Geribaldi, F. Guittard,
The effect of the inclusion of salt on the aggregation behavior of the Polymerizable semi-fluorinated gemini surfactants designed for antimicrobial
aqueous solutions of multiheaded surfactants has been studied using all- materials, J. Colloid Interface Sci. 332 (1) (2009) 201–207.
[5] P. Gilbert, L. Moore, Cationic antiseptics: diversity of action under a common
atom molecular dynamics simulations. KBr and NaSal salts were added
epithet, J. Appl. Microbiol. 99 (4) (2005) 703–715.
at three different concentrations to investigate the effect of salt con­ [6] E. Berggren, M. Liljedahl, B. Winbladh, B. Andreasson, T. Curstedt, B. Robertson,
centration on the structure of the solution. Aggregates were observed to J. Schollin, Pilot study of nebulized surfactant therapy for neonatal respiratory
distress syndrome, Acta Paediatr. 89 (4) (2000) 460–464.
form in the solution by the association of cations and were stable
[7] S. Paria, Surfactant-enhanced remediation of organic contaminated soil and water,
throughout the simulation. Adv. Colloid Interface Sci. 138 (1) (2008) 24–58.
Increase in the number of head groups reduces the organization of [8] J. Eastoe, P. Rogueda, B.J. Harrison, A.M. Howe, A.R. Pitt, Properties of a
cations around themselves due to the increased electrostatic repulsions dichained “sugar surfactant”, Langmuir 10 (12) (1994) 4429–4433.
[9] S. Bhattacharya, M. Subramanian, Synthesis of novel phosphatidylcholine lipids
between the positively charged head groups. Salicylate ions are more with fatty acid chains bearing aromatic units. generation of oxidatively stable, fluid
likely to interact with the cations, unlike the bromide ions. The salicy­ phospholipid membranes, Tetrahedron Lett. 43 (23) (2002) 4203–4206.
late ions have negligible interaction with water molecules, whereas the [10] C. Domb, Phase Transitions and Critical Phenomena, Elsevier, 2000.
[11] S. Gupta, P. Biswas, Conformational properties of complexes of poly (propylene
bromide ions have high probability to have water molecules around imine) dendrimers with linear polyelectrolytes in dilute solutions, J. Chem. Phys.
them. 153 (19) (2020) 194902.
The probability of water being around the cation increases with the [12] Y. Yang, A.K. Narayanan Nair, S. Sun, Sorption and diffusion of methane and
carbon dioxide in amorphous poly (alkyl acrylates): a molecular simulation study,
number of head groups in the cation. The increase in concentration of J. Phys. Chem. B 124 (7) (2020) 1301–1310.
the NaSal salt decreases the water density around the cations due to [13] S. De, V.K. Aswal, P.S. Goyal, S. Bhattacharya, Role of spacer chain length in
shielding of cations from water molecules. Hydration index also shows dimeric micellar organization. small angle neutron scattering and fluorescence
studies, J. Phys. Chem. 100 (28) (1996) 11664–11671.
that the NaSal salt reduces the exposure of cations to water thus pro­
[14] F.M. Menger, J.S. Keiper, Gemini surfactants, Angew. Chem. 39 (11) (2000)
moting the association of cations to form aggregates. The effect of KBr 1906–1920.
salt on promoting the aggregation is not that pronounced. [15] J. Haldar, V.K. Aswal, P.S. Goyal, S. Bhattacharya, Unusual micellar properties of
multiheaded cationic surfactants in the presence of strong charge neutralizing salts,
The observations from the RDFs and hydration index are corrobo­
J. Colloid Interface Sci. 282 (1) (2005) 156–161.
rated by the aggregation numbers in various solutions. NaSal is found to [16] T. Shikata, H. Hirata, T. Kotaka, Micelle formation of detergent molecules in
be more effective in increasing the size of the aggregates in the solution. aqueous media: viscoelastic properties of aqueous cetyltrimethylammonium
The size of the aggregates is found to decrease with the increase in the bromide solutions, Langmuir 3 (6) (1987) 1081–1086.
[17] P. Ekwall, L. Mandell, K. Fontell, The cetyltrimethylammonium bromide-hexanol-
number of head groups. The addition of NaSal increases the size of the water system, J. Colloid Interface Sci. 29 (4) (1969) 639–646.
aggregates significantly even in 3H and 4H systems. Based on the [18] S. Padsala, N. Dharaiya, N.V. Sastry, V.K. Aswal, P. Bahadur, Microstructural
analysis of the structure of the solution and the aggregation properties, it morphologies of ctab micelles modulated by aromatic acids, RSC Adv. 6 (107)
(2016) 105035–105045.
is evident that the NaSal is better promoter of aggregation compared to [19] C. Chadha, G. Singh, G. Singh, H. Kumar, T.S. Kang, Modulating the mixed
KBr salt. micellization of ctab and an ionic liquid 1-hexadecyl-3-methylimidazollium
bromide via varying physical states of ionic liquid, RSC Adv. 6 (44) (2016)
38238–38251.
Declaration of competing interest [20] C.N. Lam, C. Do, Y. Wang, G.-R. Huang, W.-R. Chen, Structural properties of the
evolution of ctab/nasal micelles investigated by sans and rheometry, Phys. Chem.
The authors declare that they have no known competing financial Chem. Phys. 21 (33) (2019) 18346–18351.
[21] S. Illa-Tuset, D.C. Malaspina, J. Faraudo, Coarse-grained molecular dynamics
interests or personal relationships that could have appeared to influence simulation of the interface behaviour and self-assembly of ctab cationic surfactants,
the work reported in this paper. Phys. Chem. Chem. Phys. 20 (41) (2018) 26422–26430.
[22] R. Makhloufi, E. Hirsch, S. Candau, W. Binana-Limbele, R. Zana, Fluorescence
quenching and elastic and quasi-elastic light scattering studies of elongated
Acknowledgments
micelles in solutions of cetyltrimethylammonium chloride in the presence of
sodium salicylate, J. Phys. Chem. 93 (24) (1989) 8095–8101.
The authors gratefully acknowledge NISER - Bhubaneswar for [23] R. Oda, I. Huc, J.-C. Homo, B. Heinrich, M. Schmutz, S. Candau, Elongated
aggregates formed by cationic gemini surfactants, Langmuir 15 (7) (1999)
providing the computational resources. S.D and U.D.C thanks Dr. Anir­
2384–2390.
ban Sharma for helping with the aggregation number code. [24] T. Shikata, H. Hirata, T. Kotaka, Micelle formation of detergent molecules in
aqueous media. 3. viscoelastic properties of aqueous cetyltrimethylammonium
Appendix A. Supplementary data bromide-salicylic acid solutions, Langmuir 5 (2) (1989) 398–405.
[25] T. Clausen, P. Vinson, J. Minter, H. Davis, Y. Talmon, W. Miller, Viscoelastic
micellar solutions: microscopy and rheology, J. Phys. Chem. 96 (1) (1992)
Supplementary data to this article can be found online at https://doi. 474–484.

9
S. Dash et al. Journal of Molecular Graphics and Modelling 111 (2022) 108110

[26] H. Hirata, Y. Sakaiguchi, J. Akai, Electron diffraction observed in the gigantic [44] S.K. Samanta, S. Bhattacharya, P.K. Maiti, Coarse-grained molecular dynamics
micelle-producing system of ctab-aromatic additives, J. Colloid Interface Sci. 127 simulation of the aggregation properties of multiheaded cationic surfactants in
(2) (1989) 589–591. water, J. Phys. Chem. B 113 (41) (2009) 13545–13550.
[27] F. Quirion, L.J. Magid, Growth and counterion binding of [45] D. Yu, X. Huang, M. Deng, Y. Lin, L. Jiang, J. Huang, Y. Wang, Effects of inorganic
cetyltrimethylammonium bromide aggregates at 25 degree c: a neutron and light and organic salts on aggregation behavior of cationic gemini surfactants, J. Phys.
scattering study, J. Phys. Chem. B 90 (21) (1986) 5435–5441. Chem. B 114 (46) (2010) 14955–14964.
[28] J. Haldar, V.K. Aswal, P.S. Goyal, S. Bhattacharya, Role of incorporation of [46] L. Wattebled, A. Laschewsky, Effects of organic salt additives on the behavior of
multiple headgroups in cationic surfactants in determining micellar properties. dimeric (“gemini”) surfactants in aqueous solution, Langmuir 23 (20) (2007)
small-angle-neutron-scattering and fluorescence studies, J. Phys. Chem. B 105 (51) 10044–10052.
(2001) 12803–12808. [47] J.-H. Mu, G.-Z. Li, X.-L. Jia, H.-X. Wang, G.-Y. Zhang, Rheological properties and
[29] S. Das, F. Khabaz, Q. Nguyen, R.T. Bonnecaze, Molecular dynamics simulations of microstructures of anionic micellar solutions in the presence of different inorganic
aqueous nonionic surfactants on a carbonate surface, J. Phys. Chem. B 124 (37) salts, J. Phys. Chem. B 106 (44) (2002) 11685–11693.
(2020) 8158–8166. [48] P. Hassan, J. Yakhmi, Growth of cationic micelles in the presence of organic
[30] J.A. da Silva, M.R. Meneghetti, P.A. Netz, Molecular dynamics simulations of the additives, Langmuir 16 (18) (2000) 7187–7191.
structural arrangement and density of alkylamine surfactants on copper surfaces: [49] K. Bijma, J.B. Engberts, Effect of counterions on properties of micelles formed by
implications for anisotropic growth of copper nanowires, ACS Appl. Nano Mater. 3 alkylpyridinium surfactants. 1. conductometry and 1h-nmr chemical shifts,
(6) (2020) 5343–5350. Langmuir 13 (18) (1997) 4843–4849.
[31] C. Xu, H. Wang, D. Wang, X. Zhu, Y. Zhu, X. Bai, Q. Yang, Improvement of foaming [50] P. Mukerjee, C.C. Chan, Effects of high salt concentrations on the micellization of
ability of surfactant solutions by water-soluble polymers: experiment and octyl glucoside: salting-out of monomers and electrolyte effects on the micelle-
molecular dynamics simulation, Polymers 12 (3) (2020) 571. water interfacial tension1, Langmuir 18 (14) (2002) 5375–5381.
[32] J.C. Shelley, M.Y. Shelley, Computer simulation of surfactant solutions, Curr. Opin. [51] D. Van Der Spoel, E. Lindahl, B. Hess, G. Groenhof, A.E. Mark, H.J. Berendsen,
Colloid Interface Sci. 5 (1–2) (2000) 101–110. Gromacs: fast, flexible, and free, J. Comput. Chem. 26 (16) (2005) 1701–1718.
[33] R. Rajagopalan, Simulations of self-assembling systems, Curr. Opin. Colloid [52] W.L. Jorgensen, D.S. Maxwell, J. Tirado-Rives, Development and testing of the opls
Interface Sci. 6 (4) (2001) 357–365. all-atom force field on conformational energetics and properties of organic liquids,
[34] S. Burov, N. Obrezkov, A. Vanin, E. Piotrovskaya, Molecular dynamic simulation of J. Am. Chem. Soc. 118 (45) (1996) 11225–11236.
micellar solutions: a coarse-grain model, Colloid J. 70 (1) (2008) 1–5. [53] H. Berendsen, J. Grigera, T. Straatsma, The missing term in effective pair
[35] J. Wendoloski, S. Kimatian, C. Schutt, F. Salemme, Molecular dynamics simulation potentials, J. Phys. Chem. 91 (24) (1987) 6269–6271.
of a phospholipid micelle, Science 243 (4891) (1989) 636–638. [54] B. Hess, H. Bekker, H.J. Berendsen, J.G. Fraaije, Lincs: a linear constraint solver for
[36] B. Jönsson, O. Edholm, O. Teleman, Molecular dynamics simulations of a sodium molecular simulations, J. Comput. Chem. 18 (12) (1997) 1463–1472.
octanoate micelle in aqueous solution, J. Chem. Phys. 85 (4) (1986) 2259–2271. [55] W.G. Hoover, Canonical dynamics: equilibrium phase-space distributions, Phys.
[37] J. Boecker, J. Brickmann, P. Bopp, Molecular dynamics simulation study of an n- Rev. A 31 (1985) 1695–1697.
decyltrimethylammonium chloride micelle in water, J. Phys. Chem. 98 (2) (1994) [56] M. Parrinello, A. Rahman, Polymorphic transitions in single crystals: a new
712–717. molecular dynamics method, J. Appl. Phys. 52 (12) (1981) 7182–7190.
[38] J.-B. Maillet, V. Lachet, P.V. Coveney, Large scale molecular dynamics simulation [57] T. Darden, D. York, L. Pedersen, Particle mesh ewald: an n-log (n) method for
of self-assembly processes in short and long chain cationic surfactants, Phys. Chem. ewald sums in large systems, J. Chem. Phys. 98 (12) (1993) 10089–10092.
Chem. Phys. 1 (23) (1999) 5277–5290. [58] W. Humphrey, A. Dalke, K. Schulten, Vmd: visual molecular dynamics, J. Mol.
[39] P.K. Maiti, D. Chowdhury, Micellar aggregates of gemini surfactants: Monte Carlo Graph. 14 (1) (1996) 33–38.
simulation of a microscopic model, Europhys. Lett. 41 (2) (1998) 183–188. [59] W.L. DeLano, Pymol: an open-source molecular graphics tool, CCP4 Newsletter on
[40] M.L. Klein, W. Shinoda, Large-scale molecular dynamics simulations of self- protein crystallography 40 (1) (2002) 82–92.
assembling systems, Science 321 (5890) (2008) 798–800. [60] N. Michaud-Agrawal, E.J. Denning, T.B. Woolf, O. Beckstein, Mdanalysis: a toolkit
[41] S. Bandyopadhyay, M. Tarek, M.L. Lynch, M.L. Klein, Molecular dynamics study of for the analysis of molecular dynamics simulations, J. Comput. Chem. 32 (10)
the poly (oxyethylene) surfactant c12e2 and water, Langmuir 16 (3) (2000) (2011) 2319–2327.
942–946. [61] G.R. Desiraju, T. Steiner, The Weak Hydrogen Bond: in Structural Chemistry and
[42] S. Abel, F. Sterpone, S. Bandyopadhyay, M. Marchi, Molecular modeling and Biology, vol. 9, Oxford University Press, 2001.
simulations of aot- water reverse micelles in isooctane: structural and dynamic [62] H.R. Bosshard, D.N. Marti, I. Jelesarov, Protein stabilization by salt bridges:
properties, J. Phys. Chem. B 108 (50) (2004) 19458–19466. concepts, experimental approaches and clarification of some misunderstandings,
[43] A. Sharma, B. Bhargava, Self-assembly of cations in aqueous solutions of J. Mol. Recogn. 17 (1) (2004) 1–16.
multiheaded cationic surfactants: all atom molecular dynamics simulation studies, [63] J.E. Donald, D.W. Kulp, W.F. DeGrado, Salt bridges: geometrically specific,
J. Phys. Chem. B 122 (48) (2018) 10943–10952. designable interactions, Proteins 79 (3) (2011) 898–915.
[64] S. Pylaeva, M. Brehm, D. Sebastiani, Salt bridge in aqueous solution: strong
structural motifs but weak enthalpic effect, Sci. Rep. 8 (1) (2018) 1–7.

10

You might also like