You are on page 1of 23

Introduction to Surfactant Self-Assembly

Laurence S. Romsted
Rutgers, The State University of New Jersey, New Brunswick, NJ, USA

1 Introduction
2 Surfactants and Their Self-assembly
3 Characteristics of Self-assembled Surfactant
Solutions
4 Modeling the Properties of Surfactant Solutions
5 Forces Contributing to Surfactant Aggregate
Structures
6 Association Colloid Effects on Chemical
Reactivity, an Introdution
7 Conclusions
Acknowledgments
References
Further Reading

1
3
5
8
11
15
20
20
20
22

INTRODUCTION
If there is Magic on the Planet, It is Contained in Water.
(Loren Eisley, The Immense Journey, 1957)

The spontaneous self-assembly of surfactants is an active


area of research in part because weak interactions in solutions of ionic surfactants depend on both ion type and
charge and these ion specific effects alter, sometimes dramatically, the chemical and physical properties of surfactant
solutions. However, consensus is absent on how to model
these effects. In modern science terms this is an ancient
problem because specific ion effects of surfactants solutions, proteins and biomembranes have been known for

more than a century.13 But, this problem is an infant in


terms of the preparation and use of soap by human societies. Soap making was probably first described in writing
by the Babylonians in about 3000 BC.4 The basic process
is the saponification or base catalyzed hydrolysis of fats
or oils from flora or fauna (Scheme 1). In early times, the
base came from wood ashes or lye (contains NaOH) and the
hydrolysis produced glycerol and the sodium salts of mixtures of long chain, saturated and unsaturated carboxylic or
fatty acids, RCO2 Na+ . In more recent times, the methods
for producing soaps and other types of surface active agents,
variously called surfactants, amphiphiles, detergents, and
soaps, have been refined and enormously expanded. However, perhaps the most profound social change is the virtually universal use of soap among human societies for
hygiene with substantial reductions in the levels of nasty
microorganisms and, serendipitously, unappreciated odors
among people.4 Special microemulsion formulations are
also used as gentle cleaning solutions for paintings and
frescos.5
In the second decade of the last century, J. W.
McBain proposed that long-chained alkyl carboxylates form
micelles: The novel suggestion is advanced that it may
be due to highly charged aggregates or micelles exhibiting
even an equivalent conductivity comparable with that of
ordinary ions.6 The unusual properties of soaps in water,
for example, conductivity, but also detergency, surface
tension reduction, and solubilization of dyes (and greasy
dirt), above a certain concentration, now called the critical
micelle concentration (cmc), are caused by the formation
of aggregates of soap molecules (Figure 1).7
Below the cmc, surface-active molecules are generally
distributed in solution as isolated molecules called unimers
(or monomers), and, despite initial scoffing, McBains basic
proposal, the formation of micellar aggregates above cmc
has stood the test of time and remains the basic concept

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Concepts
O
R

O
O

+ 3 NaOH

H 2O

HO

O
3R

O Na+

HO
HO

Sodium
alkylcarboxylates

Triglyceride

Glycerol

Scheme 1 Preparation of soap begins with the saponification or


basic hydrolysis in NaOH of triglycerides, triesters of glycerol.
The alkyl groups of the esters, R, are mixtures of saturated chains
and unsaturated chains with one or more cis double bonds and
various chain lengths. The average chain length, that is, the degree
of unsaturation, depends on the source.

Air
Water

Nonpolar
tail

mixtures of amphiphiles of various types are the structural


elements of biological membrane bilayers of cells and cell
compartments, for example, mitochondria, chloroplasts, and
vacuoles.9 Often the terms amphiphile or surfactant can be
used interchangeably and the word surfactant is generally
used here.
What keeps the field of surfactant self-assembly vibrant
and expanding is the exquisite sensitivity of the surfactant
aggregate structure to seemingly small changes in noncovalent, weak (circa 20 kJ mol1 ) interactions that depend
not only on tail length, headgroup structure, and counterion type, but also on surfactant and salt concentration, the
presence of organic additives, temperature, and the plethora
of methods required to understand surfactant solution properties.10 Thus, the aggregate properties of surfactant solutions sometimes respond dramatically to small changes in
the surfactant structure and solution composition. Different
aggregates have a wide variation in their properties that are
important for current and potential applications in many
fields as described in encyclopedias11, 12 and in the Surfactant Science Series, edited books on a variety of specialized
topics.13

1.1
Unimers

Polar
headgroup

Micelle

Figure 1 Cartoon of the surfactant unimer distribution in aqueous solution and the spontaneous self-assembly of surfactant
unimers into spherical micelles just above the cmc. The two
sets of arrows represent the concept of dynamic equilibrium in
which the exchange rates of unimer between the air/water interface and micelles are equal. The cross section of the spherical
micelle shows the core region containing the tails, the interfacial region containing hydrated headgroups (and a fraction of the
counterions for ionic micelles, not shown), and the surrounding
aqueous region. Such iconic images of micelles are unrealistic
because experiments show that micelles are fluids and the tails
are almost randomly distributed, and headgroups move at near
diffusion-controlled rates that do not define a smooth surface.

for interpreting the properties of soap solutions.8 Surfaceactive molecules reduce the surface tension at the airwater
interface, reduce the curvature of water drops on nonpolar surfaces, enhances foaming, and makes water feel
slippery or soapy. Detergents, natural or synthetic, are
surface-active molecules specially crafted for washing efficiency. The term surfactant (surface active agent) generally refers to synthetic compounds of high purity (commercial or prepared in the laboratory) with structures
designed for particular research applications. Amphiphiles
are surface-active molecules found in living systems, and

Scope of the chapter

The primary aim is to introduce the current concepts used


to interpret the properties of homogeneous, optically transparent, self-assembling aqueous solutions of small molecule
surfactants that form into association colloids composed of
charged or uncharged surfactants into micelles, microemulsions, vesicles, or other mesophases. Pseudophase models are used to interpret chemical reactivity in surfactant
solutions. Large surface-active molecules such as proteins,
starches, and polymers are not considered. Much of the
information is on surfactant solutions at room temperature and atmospheric pressure because most of the important properties, concepts, and unanswered questions can
be developed at ambient conditions. Effects of additives
such as salts, alcohols, and oils, and temperature are introduced briefly. Many introductory books include substantial
sections on surfactant self-assembly.2, 1418 Current research
on a variety of topics is periodically reviewed in Current
Opinion in Colloid and Interface Science.

1.1.1 Brief background on the balance-of-forces


concept
Tanford developed the idea that micelle structure and surfactant solution properties depend on a delicate balance
of forces, that is, the hydrophobic effect drives aggregation and is balanced by stabilizing interactions in the
interfacial region of the aggregates between headgroups

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly

= Surfactant. Alkyl tail typically 818 carbons straight chain


Polar head, nonionic, zwitterionic, cationic, anionic

and water.19 This concept was enriched and expanded


by several groups: Nagarajan,2, 20 Ruckenstein,21, 22 and
Blankschtein.2325 Israelachvili, Mitchell, and Ninham14, 26
proposed the molecular-packing-parameter concept to correlate the shape of the surfactant and aggregate.

1.1.2 Focus on solution properties and their


relationship to the surfactant structure
Workers in the colloid and surface science field strive constantly to relate the surfactant structure to the structures
and properties of their aggregates produced by spontaneous
self-assembly. This is a complex, difficult problem because
the structures of fluid, surfactant aggregates are governed
by small changes in multiple noncovalent interactions, for
example, coulombic, hydrogen bonding, dipole or hydration that are responsible for apparently large changes in the
size and shape of the aggregate. Most models are qualitative and function more as guides to thinking about the
relationship between the surfactant tail length and structure, headgroup structure and charge, and counterion type.
Surfactant solution properties are generally described either
by the mass action or more commonly used pseudophase
models. Other important general concepts include cooperativity, the hydrophiliclipophilic balance (HLB) of the
surfactant, the packing parameter, electrolyte screening,
interfacial hydration, and ion-specific interactions. These
concepts and the intermolecular and ionic interactions governing them are difficult to characterize quantitatively, especially over a range of experimental conditions, for example,
surfactant concentration, temperature, and salt and organic
additive effects. Other important examples of self-assembly
and surfactant-like organized structures are as follows: selfassembly of structures,27, 28 DNA interactions with surfactants and polymers,29 block copolymers and nanoscience,30
dendrimers (single-molecule micelles),31 and monolayers.32, 33

2.1

SURFACTANTS AND THEIR


SELF-ASSEMBLY
What are surfactants?

Surfactants have two basic structural components, a water


insoluble, hydrophobic part usually called a tail, attached to
a water-soluble headgroup as illustrated by the

spermatozoa-like cartoon above.1517 In surfactant chemistry, cartoons are ubiquitous. They capture the essence
of structure, but often totally misrepresent the details.
Scheme 2 shows a number of examples of surfactants, both
typical and a few atypical.34
Many surfactant tails are single chains, but may have
two or even three chains, branched chains, chains with
double bounds or even triple bonds, or even containing
aromatic groups. Tails with fewer than eight carbons may
never reach sufficient concentrations to aggregate. Tails of
18 carbons or greater have a high probability of being
insoluble. Gemini surfactants have two tails attached to
two headgroups linked with spacer groups of various sizes.
They are currently an active area of research.35, 36 Surfactant
tails are attached to highly water-soluble polar or charged
headgroups, and the headgroup structure also determines
their type: cationic, anionic, nonionic, zwitterionic, and
sometimes amphoteric unimers.
Ionic surfactants have many types of headgroups and
counterions. Cationic surfactants have positively charged
headgroups and anionic surfactants have negatively charged
headgroups with their respective oppositely charged counterions. Headgroups and counterions are commonly monovalent, but both may also be di or even trivalent. Some counterions precipitate surfactants, for example, added Ca2+
precipitates long-tail alkyl carboxylates (creating the infamous bathtub scum), and the added ClO4 often precipitates quaternary ammonium surfactants, but, and no one
knows why, not zwitterionic sulfobetaines.37 Zwitterionic
surfactants contain equal numbers of positive and negative
charges in the same headgroup, but can bind ions selectively. The most important zwitterionic surfactants are biological phospholipids.38 Their cationic and anionic groups
may be in either order with various spacer lengths between
them (Scheme 2). Amphoteric surfactants have titratable
headgroups (amine, carboxylic acids, or amine oxides) and
their micelle surface charge depends on solution pH. For
example, addition of acid to micelles of an amine oxide,
Scheme 2, protonates the oxygen (pKa 4.5) and the headgroups at the micellar interface change from neutral to
cationic. Nonionic surfactant headgroups include oligooxy
chains or sugars.
Surfactants, both ionic and neutral, are also commonly
characterized qualitatively by their HLB values,15, 17 related
to the hydrophobicity of their tails that increases with the
alkyl chain length or the number of carbons in the tail versus
the polarities or water solublities of the various headgroup
and counterion types. The hydrophilicities of ionic surfactants with different counterions are difficult to compare

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Concepts

Anionic surfactants

Cationic surfactants
+

R-OSO3

M+

R-OPO3H

Alkyl sulfate,

M+

R-CO2

Alkyl phosphate

M+

R N

Alkyl carboxylate

R-N(CH3)3

Alkyltrimethylammonium

Alkylpyridinium

M = Li+, Na+, K+, Rb+,

SO3 Na+

X = F, Cl, Br, I, NO3, RCO2, SO42, tosylate,


benzoates, CH3SO3, and so on.

Dialkyl sulfosuccinate, AOT


Zwitterionic surfactants

O
Nonionic surfactants

OH

R-(OCH2CH2)n-OH

n, Wide variation

Alkyl ethoxylate

OH
OH

HO

Alkyl glucoside

CH3

R
R

R N CH2CH2CH2SO3

CH3
Alkyl sulfobetaine

+(H

Others
CO2 Na+

OH

2+
N

HO

Bile salt, sodium cholate

HO

O
O

B block nonionic surfactant


E6C10E6C10E6

Alkyl amine oxide


amphoteric

R = C14H29
HO

CH3

OH
H

A gemini, Cu catalyst

O O
O P O

CH3
+
R N O

2Cl

3C)3N

Phosphatidyl choline
R, R may be unsaturated

Cu

X
+
N

R
Gemini (n 1)

Ca2+, Mg2+, and so on.

X
+
N
R

O
O
O

O
O

O
O

O
O

O
O

O
O

Scheme 2

Some representative and some unconventional surfactant structures, A and B, Ref. 34.

and they depend on salt concentration and type. Many


properties of ionic surfactants, for example, cmc, aggregation number, and their phase diagrams, depend on the headgroup structure and counterion type. There is no consensus
on how to model such ion-specific effects. Single-chain
nonionic surfactants such as the alkyl ethoxylates, Cn Em ,
whose properties vary systematically with chain length and
oligooxy headgroup length, have the most meaningful HLB
numbers.15
Salts of bile acids (bile salts) are important in digestion by vertebrates, people included, and are also present
in potentially pain-producing gallstones. The self-assembly
of bile salts produces aggregates that are distinct and different from other single-chain surfactants because of their
rigid steroid ring structures, Scheme 2, and the asymmetric
distribution of hydroxyl groups on the fused rings makes

one side of the molecule more polar than the other more
hydrophobic side. Additional information on the formation
and properties of bile salt micelles, the cmcs, aggregation
numbers, shapes, and counterion binding is given in a recent
review.39

2.2

What is surfactant self-assembly?

Surfactant self-assembly in solution is a spontaneous


stochastic (or random) process driven by highly cooperative,
noncovalent interactions and composed of reversible steps
that produce the same set of surfactant structures at equilibrium. That is, surfactant self-assembly occurs spontaneously without human intervention beyond simple mixing.
At thermodynamic equilibrium, each solution preparation

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly

CHARACTERISTICS OF
SELF-ASSEMBLED SURFACTANT
SOLUTIONS

cmc

Osmotic pressure
Detergency

Solution property

of the same stoichiometric composition contains the same


concentrations of unimers and aggregates with the same
distribution aggregate sizes and shapes. Self-assembly also
characterizes the organization of surfactant monolayer solid
or liquid surfaces, including interfacial domains such as the
organization of biomembrane surfaces,40 and the formation
of nanocrystals, the self-assembly of cage molecules, and
the assembly of proteins into larger structures as observed
in the tobacco mosaic virus.28

NMR chemical shift

Stabilization

NMR selfdiffusion

Turbidity

Surface tension

3.1

The critical micelle concentration (cmc)

Any measurable difference in a surfactant solution property in the absence and presence of micelles, that is, below
and above the cmc, can be used to determine the cmc,
which is one of the most important properties of association
colloids.15, 17, 18, 41 Figure 2 illustrates typical changes in a
number of physical properties with increasing surfactant
concentration from below to above the cmc. Each method
shows a break in the surfactant-concentration-property profile consistent with micelle formation. For example, solubility and turbidity increase because micelles dissolve
water-insoluble molecules and scatter UV light. The NMR
self-diffusion of a surfactant molecule decreases above the
cmc because micellized unimers diffuse more slowly. 1 H
and 13 C NMR signals shift as the unimer moves from an
aqueous to micellar environment, which has a lower polarity than water. The conductivity and the electrical potentials
of surfactant solutions decrease above the cmc for ionic surfactants because the ions associate with micellar surfaces
and do not contribute to the measured conductivity or potential. Below the cmc, adding the surfactant increases the
unimer concentration in the aqueous solution and also in the
interfacial region (Figure 1). Increasing the concentration of
the surfactant at the air/water interface decreases the surface
tension because the surfactant tails associate with the lower
polarity environment of air. The increase in concentration
of surfactant alkyl tails lowers the surface tension because
the intermolecular forces between surfactant molecules are
weaker than those between water molecules. Above the
cmc, virtually all the added surfactant makes micelles and
the unimer concentration in the water and at the interface
remain approximately constant. Surfactant molecules also
organize spontaneously at the interface between oil and
water phases and lower the interfacial tension.
The cmc is not a phase boundary because solutions
containing micelles are homogenous. But the cmc does

Interfacial tension

Equivalent
conductivity

Surfactant concentration

Figure 2 Some methods for determining the critical micelle


concentration (cmc). Bulk solution property methods that respond
to changes in some property throughout the solution include
the following: surface tension (airwater), osmotic pressure,
conductivity, turbidity (e.g., static or dynamic light scattering),
and dye solublization. Probe methods, in which the property
of some molecule or molecules show a change because of a
transition from the probes properties in water and micelles
include the following: 1 H and 13 C NMR chemical shift or change
in self-diffusion coefficient of the unimer, indicator binding
by UVvisible/fluorescence spectral shift, and changes in the
observed rate constant of a reaction.

mark a relatively rapid increase in micelle formation


over a relatively narrow concentration range from a solution composed of surfactant unimers to one composed
of aggregates.15, 18 Cmc values are often method dependent, but for pure surfactants the spread in cmc values
is generally narrow. For example, the dissolution of a
water-insoluble dye may occur at a lower concentration
than a change in 1 H NMR chemical shift because the
dye induces micelle formation at a lower concentration.42
Alternatively, the dye might induce the formation of premicellar aggregates when the dye/surfactant concentrations
are similar, but below the cmc. Cmc measurements are
also often sensitive indicators of surfactant purity. For
example, small amounts of hydrophobic impurities often
produce minima in surface tensionlog (surfactant concentration) plots.15 Also, the estimated cmc values are
generally lower than literature values for pure surfactants.
Many of these methods also provide information on micelle
properties.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

3.2

Concepts

Correlations of the cmc with surfactant


properties

are added to solutions of ionic surfactants, Scheme 2, the


extent of the decrease in the cmc is different, showing that
both the salt concentration and the ion-specific effects are
influencing the cmc.17

3.2.1 Surfactant tail length


The cmc values of pure surfactants are sensitive to the
structure of the surfactant, that is, the nature of the alkyl tail,
headgroup structure, and counterion type, the composition
of the solution, and to changes in temperature.15, 17 One
commonly observed correlation is the linear relationship
between the between log cmc and the number of carbons,
n, in the surfactant tail:
log cmc = A Bn

(1)

where A and B are empirical parameters for the intercept


and slope respectively. Adding methylenes incrementally
has an incremental effect on log cmc, consistent with
constant decrease in the free energy of transfer of a
methylene of a surfactant tail from water to the micellar
core and an incremental increase in the hydrophobic effect
that drives micellization (see below).

3.2.2 Temperature
The cmc values of nonionic surfactants decrease with
increasing temperature up to the cloud point or lower
consolute boundary at which the transparent solution turns
cloudy.15, 17 Above the cloud point, two phases are present,
one dispersed in the other. On cooling below the cloud
point, the dispersion turns transparent. This reversible
process is consistent with a decrease in the free energy of
hydration of the oligooxyethylene15 or sugar headgroups43
on heating (Scheme 2). Ionic surfactants, on the contrary,
sometimes show minima in their cmc values with increasing
temperature in the vicinity of room temperature.17 At higher
temperatures, surfactant solubility generally continues to
increase, although some surfactants with large headgroups44
or large counterions45 are exceptions.

3.2.3 Additives
Three types of additives are routinely added to micellar
solutions: (i) salts that often contain the counterion of the
surfactant or sometimes salts containing aryl rings such
as benzoates or arylsulfonates; (ii) alcohols of various
chain lengths; or (iii) a second surfactant producing mixed
micelles.
Salts.
Addition of virtually any salt reduces the cmc of
ionic surfactants significantly, by a factor of 10 or more
when the salt concentration exceeds 10100 mM, but the
effects of salt on the cmc values of nonionic surfactants
are more modest.15 When salts having different counterions

Organic additives.
Addition of alcohols (ROH) to surfactants with any type of headgroup generally reduces the
cmc when the length of the alcohol tail exceeds three carbons, that is, n-butanols and longer.46 The cmc values
decrease because ROH has only one polar functionality,
the -OH group, which makes micelles composed of surfactant and ROH more hydrophobic. Longer alcohol chains
are even more nonpolar and they induce micellization more
effectively as the alcohol chain length increases. Shortchain, water-miscible alcohols, on the contrary, tend to
increase the cmc in the order CH3 OH>EtOH>n-PrOH and
the micelles break up completely at high alcohol concentrations, for example, at or above about 50% alcohol by
volume.
Mixed micelles.
An additive may also be a surfactant.47, 48 The cmc values of mixtures of two or more surfactants depend strongly on the structural similarities of the
two surfactants.15, 48 For surfactants of similar tail lengths
and headgroups, for example, sodium dodecyl sulfate and
sodium dodecylsulfonate, the mixture behaves ideally at all
mole fractions of the two surfactants, that is, the cmc is
a simple function of the surfactant mole fraction. When
the surfactants are significantly different in terms of chain
length or headgroup structure (e.g., opposite charge), the
mixtures do not behave ideally and the cmc values require
an empirical interaction parameter, , to fit the change in
cmc with surfactant mole fraction. For example, in mixtures of anionic and cationic surfactants called catanionic
aggregates, cationanion headgroup interactions stabilize
the micelles and reduce the cmc at intermediate mole fractions compared to the cmcs of the pure components. These
systems may form single-walled vesicles spontaneously
and they are being studied for binding of DNA and other
solutes.49 Mixtures of twin tail surfactants, for example,
phospholipids, often form enriched regions in vesicles or
membrane bilayers called domains or rafts that have a
higher fraction of one surfactant over another.40

3.3

Phase diagrams

Although much surfactant-based research has focused on


dilute aqueous solutions of surfactant micelles, they are
only a small fraction of the total composition range that
may be studied. Phase diagrams have been prepared for
a wide range of surfactant systems containing two or
three components, binary and ternary phase diagrams,

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly

Surfactant

Cubic

Vesicle

Cubic
Lamellar

Hexagonal
Irregular
bicontinuous

Cylindrical

Spherical Water
micelles
(Oil-in-water)

Oil

Inverted
micelles
(Water-in-oil)

Figure 3 Hypothetical ternary phase diagram for a three-component system of water, oil, and surfactant showing the shapes of
various aggregate structures. Vesicle bilayers, which are often large, have surface curvatures approaching lamellar. The open regions
are homogeneous solutions of single mesophases. The regions marked by tie lines are biphasic and the composition of the two phases
are those defined by the contacts of the ends of the lines with homogeneous regions. The dark gray regions are triphasic. Aggregate
structures are determined by scattering methods. (Redrawn from Ref. 51. American Institute of Physics, 1989.)

respectively. Sometimes salt is included or the temperature


is varied.17, 50

3.3.1 Ternary phase diagrams


Figure 3 shows a hypothetical phase diagram for a threecomponent ternary system of water, oil, and surfactant
at constant temperature.51 The structures illustrate those
commonly observed in three-component mixtures of water,
oil, and surfactant and also in two components of mixtures of water and surfactant (see below). Details regarding
reading and usage of phase diagrams are published elsewhere.17, 50
The three sides of a ternary diagram define binary
mixtures expressed in weight percent (wt%): water and
oil, water and surfactant, and oil and surfactant. The
open areas represent homogeneous, single-phase solutions;
two-phase regions are marked with so-called tie lines;
and three-phase regions are solid gray. For example,
because oil and water do not mix, the entire length of
the wateroil side denotes two separate phases. However,
the watersurfactant and oilsurfactant sides that are
boundaries to open regions represent single phases. Any
point within the three sides denotes a water/oil/surfactant
mixture that contains a specific weight percentage of each.
The ends of each tie line are connected to a homogeneous
region whose composition defines the composition of the
two coexisting phases. The bold dashed line drawn from
the watersurfactant side at any composition, for example,

50% water and 50% surfactant, to the oil corner, shows


the consequence of adding oil at a constant weight ratio
of surfactant and water. The initial two-phase system
passes through a lamellar region, a three-phase region,
a bicontinuous region, a reverse microemulsion droplet
region, and ends in a two-phase oil/water region as the
weight percentage of water and surfactant approaches
zero. A curved path through the single-phase region that
extends from the water corner to the oil corners marks a
continuous transition from oil-in-water droplets to waterin-oil droplets with a irregular bicontinuous region in the
middle; a complete inversion of droplet structure without
phase separation.
The structures of the aggregates in the homogeneous
mesophases in the presence of oil or other additives include
the following: spherical, rodlike or cylindrical, hexagonal
close packed, cubic (two types), lamellar, irregular bicontinuous, and reverse micelles. The shapes of the structures in
each region are usually measured by scattering techniques,
for example, light (dynamic and static), small angle X ray,
and neutron scattering, and by cryo-TEM.52, 53 The irregular bicontinuous region is of particular interest because the
structures are fluid, like waving fingers of oil-in-water (or
water-in-oil) with the boundary coated by a monolayer of
the surfactant.54

3.3.2 Binary phase diagrams, ionic surfactants


Figure 4 shows the binary diagram for the quintessential
anionic surfactant, sodium dodecyl sulfate (SDS) in water

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Concepts
100

Ra Qa T a

Temperature (C)

80

Lam

X8.W

Hex

60
Mic

40

Hex+X.W2

X.W

20

X.W2

Liq+X.W2

Liq+X.W2

0
Ice+X.W2

20

20

40

60

Sodium dodecyl sulfate (%)

80
X.W2

100
X.W
X.W?

Figure 4 Binary phase diagram for sodium dodecyl sulfate as


a function of temperature. Note that SDS has a substantial region
with different hydrated solids in water and ice and that the Krafft
temperature increases with temperature. At very high wt% SDS,
solid hydrates are formed. (Redrawn from Ref. 17. John Wiley
& Sons, Ltd, 1998.)

with increasing surfactant from 0 to 100 wt% as a function temperature (20 to 100 C).The upper part of the
diagram (2080 C) shows a transition from aqueous
micelles to hexagonal to lamellar mesophases, separated
by slender, two-phase regions with increasing SDS wt%.
The line separating the micellar from the liq+X.W2 phases
up to about 40 wt% SDS marks the Krafft temperature,
the temperature above which micelles form and below
which the solution contains a mixture of hydrated SDS,
(X.W2 ), and water. The remaining regions are various
types of crystalline hydrates. Binary phase diagrams have
been created for a wide variety of surfactants17, 50 and
most show common structural transitions of the homogenous mesophases with increasing surfactant weight percentage and decreasing water weight percentage from micelles
to hexagonal to lamellar to reverse hexagonal to reverse
micelles, sometimes called the Fontell Scheme.55 However, the size and shape of the various regions is sensitive to surfactant chain length, headgroup structure, and
counterion type. Phase diagrams also have been prepared
containing additives such as salts and alcohols, leading to four component and higher phase diagrams. Their
presence affects the sizes and shapes of the mesophase
regions.50

3.3.3 Binary phase diagrams, nonionic surfactants


Similar to ionic surfactants, nonionic surfactants such as
alkyl ethoxylates, Cm En , and alkyl sugars form micelles
and microemulsions above their cmcs.15, 17, 5658 Cm En surfactants contain hydrocarbon chains of variable lengths
and the En units may vary from short, E3 , to very
long, for example, commercial Brij 35P has the formula

C12 E23 . Similar to ionic surfactants, Figure 4, nonionic surfactants have complex binary phase diagrams containing
many of the same mesophases as those in ionic surfactants.
Increasing the surfactant concentration generally results in
structural transitions such as those observed with ionic
micelles, but over different temperature and concentration ranges. Cubic phases may also be present depending
upon the length of the oligooxy chain.59 The concentration and temperature boundaries of the mesophase regions
correlate with the HLB values of the nonionic surfactants.
The cmc values of nonionic surfactants decrease more
rapidly with increasing alkyl chain length than the cmc
values of ionic surfactants. However, unlike ionic surfactants, whose solubilities increase with temperature, nonionic
micelles grow with increasing temperature until they phase
separate at the lower consolute boundary or cloud point.
For a constant alkyl tail length, the cloud point increases
rapidly with increasing oligooxy chain length. Above the
cloud point temperature, the micellar solution is biphasic
and cloudy. The reason for aggregate growth and phase
separation is that the oligooxy chains are not as strongly
hydrated as the charged headgroups, and counterions of
ionic surfactants and increase in temperature results in
weaker hydration of the interfacial region and loss of interfacial water, closer packing, increased aggregation numbers, and eventually phase separation. Measurements of the
hydration number of the headgroup region and the number
of water molecules per ethylene oxide unit, by NMR selfdiffusion60 or by the chemical trapping method,61, 62 show
these hydration decreases clearly.

MODELING THE PROPERTIES OF


SURFACTANT SOLUTIONS

Two approaches are generally used for modeling the


properties of surfactant solutions, pseudophase and the
mass action models. The pseudophase model is easier to
use and has wider applicability in experimental work. For
example, the pseudophase model is the basis for current
interpretations of association colloid effects on the rates
and equilibria of chemical reactions (Section 6).

4.1

Pseudophase model

Pseudophase models17, 48, 63 were originally developed to


treat mixed micellization of binary surfactants.48 The concept has been generalized to mean that the totality of
the surfactant aggregates present, for example, micelles,
microemulsions, vesicles, and so on, in homogeneous solutions are treated conceptually as a separate phase, or

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly

Table 1 Rate constants for exit, k , and entrance, k + , and exit rate constants for the surfactant monomer from SDS micelles,
dimyristoylphosphatidylcholine (DMPC), and pyrene from DMPC vesicles.
Surfactant
C6 H13 OSO3 Na+ a
C8 H17 OSO3 Na+ a
C12 H25 OSO3 Na+ a
C14 H29 OSO3 Na+ a
DMPC vesicleb
DMPC vesicleb
a
b

Component, conditions

Cmc (M)

107 k (s1 )

25 C
25 C
25 C
25 C
DMPC
Pyrenec

0.42
0.13
0.0082
0.00205
0
0

132
10
1
960
1000
22

108 k + (M1 s1 )
32
77
12
4.7
No value
No value

Ref. 17.
Ref. 65.
4% EtOH.

c 20 C,

pseudophase. Consequently, pseudophase models do not


include assumptions about aggregate size and shape. The
pseudophase model works for two reasons: (i) micelle
formation is an extraordinarily cooperative process such
that about 30100 or more unimers aggregate spontaneously above the cmc to form spheroidal or spherical
micelles whose properties, for example, interfacial polarity,
hydration, or degrees of ionization, are not very sensitive
to increasing aggregation numbers or changes in aggregate shape, for example, spheres to rods. (ii) The surfactant unimers enter and exit the association colloids at
near-diffusion-controlled rates (Table 1). Thus, the distributions of components in association colloid solutions are
in dynamic equilibrium once the initial mixing is complete
(see Section 6). The current theoretical model for surfactant
aggregation is that surfactants assemble in a near-diffusioncontrolled, reversible, endergonic stepwise process reaching
a critical aggregation number followed by rapid exergonic
assembly into an ensemble of micelles with a distribution
of sizes.64
The exchange rate of other components between the
droplet and the aqueous region in microemulsions is
also fast, 10 ns for medium chain length alcohols and
100 ns for the surfactants.65 However, the dynamic equilibrium assumption must be applied more carefully to other
mesophases. For example, dimyristoylphosphatidylcholine
(DMPC) molecules and pyrene have very slow exit rates
from DMPC vesicles compared to their entrance rates
(Table 1). In addition, molecules move laterally within a
bilayer leaflet and flip-flop from the inside (endo) to the
outside (exo) leaflets.65 Lateral diffusion rates are faster
above the phase transition temperature, TC , than below.65
Nevertheless, the pseudophase model can be applied to
those components whose distributions are fast enough to be
described by equilibria between the aqueous solution and
aggregates or are so slow as to be effectively stationary
after initial mixing.
In the pseudophase model, the cmc is a critical point.
Below the cmc, the stoichiometric surfactant concentration,

[ST ], equals the [unimer], which increases steadily up to


the cmc. Note: square brackets here and throughout the text
indicate the stoichoimetric concentration in moles per liter
of solution volume or molarity of the surfactant, counterion,
salt, and so on. Above the cmc, the unimer concentration
remains approximately constant and equal to the cmc, and
all the additional surfactant forms the micellar pseudophase,
[SM ] and
[ST ] = [SM ] + cmc

(2)

For example, the cmc values of many common surfactants are in the concentration range of 110 mM. Thus,
if the cmc = 1 mM and [ST ] = 5 mM, and [SM ] = 4 mM,
the solution contains approximately 30 1020 unimers
(5 mM) and 24 1020 unimers in micelles (4 mM). If the
aggregation number is 50, there are 5 1019 micelles.
If the molecular weight of the surfactant is 300, the
5 mM solution contains about 0.15 g/100 ml, or 0.15%
surfactant by weight. Much experimental work in micellar solutions is carried out at room temperature (25 C)
in 5 wt% or less surfactant concentration range, see
Figure 4 as an example. Micelle radii are approximately
equal to the length of the extended surfactant chain, for
for a 12-carbon chain and sulfate headexample, 20 A,
group. Consequently, micelles are too small to scatter visible light and their single-phase solutions are optically
transparent.
The pseudophase model only approximates reality
because the unimer concentration above the cmc is not
absolutely constant, but decreases gradually, for example,
with increasing concentration of ionic surfactants.16 However, as [ST ] increases, the fraction of the free unimer
rapidly becomes a small fraction of [ST ] (e.g., [ST ]>10
cmc), making the approximation reasonable. Also, depending on surfactant structure, small aggregates may be present
in the vicinity of the cmc that cannot be accounted for using
the pseudophase model.42 However, this problem is not
severe unless components are added that interact strongly

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

10

Concepts

or specifically with the surfactant, for example, significant


quantities of counterions or organics, or other surfactants
of opposite charge that make mixed micelles. The major
advantage of the pseudophase model is its simplicity that
permits doing experiments without measuring the size or
shape of the micelle.

4.2

The mass action model

The mass action or equilibrium model expressed as a


dynamic equilibrium of n unimers with micelles that
contain n unimers16, 17 :
nSw  Sn
[Sn ]
K=
[Sw ]n

=
(3)

Values of K can be estimated in the vicinity of the cmc


provided n is not too large. For example, this approach has
been applied to bile salt micelles that tend to form small
aggregates.39 The value of K depends on the surfactant
chain length and headgroup structure. If n is large, the
uncertainty in K grows and the conditions support using
the pseudophase model. Micellization can also be treated as
a multiple equilibrium process in which unimers get added
stepwise to growing aggregates:
S1

S1

S1

2S1  S2  S3  S4 Sn1  Sn
Kn =

[Sn ]
[Sn1 ][S1 ]

(4)

A separate equilibrium constant is written for the addition of


each unimer to the micelle. Micellization gives aggregates
with a distribution of sizes that may be broad or narrow16, 17
and Kn depends on the aggregation number such that there
is an optimum Kn value, that is, smaller values for larger
and smaller aggregates. The average aggregation number
increases steadily with surfactant concentration until it
reaches the sphere-to-rod transition concentration when
there is a dramatic increase in the aggregation number. As
noted earlier, aggregate sizes and shapes can be measured
by scattering methods and by cryo-TEM.52

4.3

At equilibrium, the micelles have a net charge that is


demonstrated by electrophoretic mobility experiments, for
example, in an electric field, cationic micelles migrate
toward a negative electrode. Typically, the surface charge
at equilibrium is about 1040% of the total number of
micellized surfactant molecules. The net surface charge
is numerically equal to the micellar degree of ionization
(or ionization degree), , which is experimentally obtained
from the measured concentration of counterions to the
aqueous pseudophase, [Xw ], corrected for the unimer concentration cmc, (5).15, 66 A variety of methods may be
used, for example, ion selective electrodes, conductivity
NMR,

The degree of ionization of ionic micelles

Both the mass action and pseudophase models are directly


applicable to nonionic micelles and zwitterionic micelles
in the absence of salt. Ionic micelles, however, have
cationic or anionic headgroups with an associated counterion, usually inorganic or organic ions that arenot sufficiently surfactant-like to drive aggregation themselves.

([Xw ] cmc)
([ST ] cmc)

(5)

self-diffusion,52 and even chemical reactions.67 In terms of


the pseudophase model, the fraction of bound counterions,
, is determined from by assuming that counterions are
either free in the aqueous pseudophase or in the micellar
pseudophase. Table 2 lists some measured cmc and values for representative anionic and cationic surfactants with
examples of both
([Xw ] cmc) ([XT ]([Xw ] cmc))
+
= + = 1 (6)
([ST ] cmc)
([ST ] cmc)
inorganic and organic counterions. Note that a number of
the values in Table 2 are between 0.2 and 0.3,15, 66 a
range that is characteristic of many inorganic counterions. Example exceptions include the following: (i) OH ,
= 0.52, a strongly hydrated, weakly polarizable ion68 ;
and (ii) tosylate ion (Tos ), an organic counterion that
associates both electrostatically and hydrophobically to the
interfacial region.69 Gemini surfactants, which have two

Table 2 Measured cmc and alpha values for selected surfactants


at 25 C unless otherwise indicated.
Surfactant
C12 H25 OSO3 Na+
C14 H29 OPO3 H Na+ a
C12 H25 N(CH3 )3 + Br
C16 H33 N(CH3 )3 + Br
C16 H33 N(CH3 )3 + OH
C16 H33 N(CH3 )3 + Tos
C16 H33 N(CH3 )3 + Cl
14-2-14 2Brb
14-2-14 2Clb

Cmc (mM)

Alpha()

8.4
16.5
65.0
0.92
3
0.26
1.32
0.14
0.26

0.22
0.29
0.26
0.16
0.52
0.13
0.37
0.23
0.26

Reference
66
66
66
15
68
69
15
70
70

a 35 C.
b
Gemini surfactants: ethanediylbis(dimethyltetradecylammonium) dibromide and dichloride.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly
counterions with each dicationic headgroup having apparently normal values.70 These observations support the
idea that ion association to interfaces is governed by both
electrostatic and specific interactions such as ion pairing
between headgroups and counterions, hydration, and polarization (see below).

FORCES CONTRIBUTING TO
SURFACTANT AGGREGATE
STRUCTURES

The sizes and shapes of surfactant aggregates is determined


by the balance of opposing forces.19 This section focuses
primarily on ionic aggregates because a greater variety
of interactions contribute to the balance of forces than in
nonionic micelles.

5.1

Balance of forces in ionic micelles

in the absence of added salt.71 Figures 5(b) and (c) are


from a recent MD simulation of decyltrimethylammonium
bromide, DeTAB, with an aggregation number of 30.72
The images in Figure 5(a) and (b) are more realistic representations of a snapshot of the distributions of components within ionic micelle than the cartoon in Figure 1.
Both Figure 5(a) and (b) reflect the disordered micelle
core and the array of headgroups and counterions in
the interfacial region spread over the micellar surface.
Figure 5(a) has minimal water/hydrocarbon contact that differs from simulation, Figure 5(b), which has substantial
water/hydrocarbon contact. Figure 5(c) shows both unitedatom and all-atom simulations of radial distribution functions for each component from the center of mass (COM).
Note the substantial overlap of tails with water in the interfacial region, 1 nm. The density of Br ions at maximum for Br , at 1.7 nm is 100 kg m3 (1.3 M), and of
quaternary ammonium headgroups, which is closer to the
COM at 1.4 nm that is about twice as large ( 2.6 M).
Results from chemical-trapping experiments report similar
estimates of interfacial counterion molarities.3 Estimates of
values are 0.103 and 0.218, respectively, for the all-atom
and united-atom models. The maxima of the headgroups
and counterions have different COMs indicating that interfacial charge density is radially polarized. The substantial
residual hydrocarbonwater contact in the interfacial region
of micelles, first proposed 30 years ago,8 is observed in
other MD simulations as well, for example, for SDS.73
Such simulations create a new image of the micellar interfacial region, compare Figure 5(a) and (b). The importance
of significant amounts hydrocarbon exposure for micelle
properties such as growth of spherical micelles and sphereto-rod transitions in the presence and absence of added salt
are not known.

Density (kg m3)

The spontaneous aggregation of ionic surfactants produces


an array of structures, whose form depends on the surfactant
structure and concentration, oil type and concentration, and
temperature (Figures 3 and 4). A completely mixed, singlephase micellar solution is at its minimum free energy. This
free energy is determined by multiple intermolecular and
ionic interactions that include electrostatic, dipoledipole,
iondipole, ionion, and dispersion (van der Waals) interactions. At equilibrium, the opposing forces are in balance
and the net force is zero. The balance determines the size
and shape of the micelle. Figure 5(a) is an older conceptual
representation of a spherical micelle in aqueous solution

(a)

(b)

11

(c)

900
800
700
600
500
400
300
200
100
0

Tails UA
Heads UA
Bromide UA
Water UA
Tails AA
Heads AA
Bromide AA
Water AA

0.5

1.5

2.5

Distance to micelle COM (nm)

Figure 5 Images of core, interfacial, and aqueous regions of ionic micelles. (a) It shows a classical micelle cartoon that represents
the dynamic nature of the course and interface, but may over emphasize the minimization of waterhydrocarbon contact. (b and c)
The results of united-atom (thin lines) and all-atom (thick lines) molecular dynamic simulations of decyltrimethylammonium bromide,
DeTABr containing 29 or 30 surfactants. COM = center of mass. (Structure 5(a): Reproduced from Ref. 71. American Chemical
Society, 1991. Structures 5(b) and 5(c): Reproduced from Ref. 72. American Chemical Society, 2008.)
Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

12

Concepts
Geometric
shape

Sphere

Cylinder/rod

Mesophase

Micelle

Hexagonal

Packing
no /ae /o
Parameter

>01/3

1/31/2

Examples

Bilayer

Sphere

Lamellar, Reverse
vesicle
micelle

1/21

SDS low
CTAB high Twin tail
concentration concentration

>1
AOT

Scheme 3 Example of relationships between the geometric


shape of an aggregate (see Figure 3), surfactant structures, shapes
of their tails relative to their headgroups of constant size, their
packing parameters, and examples of surfactants that form these
structures in some concentration range in their phase diagrams.

5.2

The packing parameter

In 1976, Israelachvili, Mitchell and Ninham proposed the


packing parameter concept and demonstrated that, at equilibrium, the aggregate size and shape can be related almost
intuitively to the surfactant structure and the thermodynamics of micellization.14, 26 For example, compare Scheme 3
and Figures 3 and 4. The cone-shaped surfactants with a
single headgroup that has a larger cross-sectional area than
the tail can be packed into spheres, whereas cylindrical
shapes with similar cross sections of tails and headgroups
pack better into rodlike aggregates. The packing parameter, P , expresses the relationship between surfactant and
aggregate shapes:
v0
P =
ae l0

(7)

where v0 and l0 are the volume and length of the surfactant


tail and ae is the surface area of the hydrophilic headgroup
per molecule in the aggregate. As the tail cross-section area
increases, the value of P increases, and the convex curvature of the surface decreases (structures in Scheme 3). For
reverse micelles and microemulsions droplets, the curvature becomes concave. Formally, the v0 / l0 ratio is constant
and independent of the tail length (the definition of v0
contains l0 ). Only ae reflects the effect of specific surfactant structures, for example, headgroup and counterion
type and concentration. However, the results in Figure 4
for SDS and the representative ternary phase diagram in
Figure 3 illustrate the fact that the P value for a particular surfactant structure in the presence and absence of
oil forms multiple aggregate structures and that a single
value of ae cannot be the sole determinant of the aggregate
shape. Indeed, many surfactant aggregates show the same
characteristic shape changes observed in Figures 3 and 4,

that is, the structural transition from convex to lamellar to


concave, with decreasing water concentration (or increasing
surfactant concentration). This decrease in surface curvature with increasing surfactant concentration suggests that,
as the amount of water in the system decreases, there should
be a concomitant decrease in the concentration of water in
the interfacial regions of the aggregates, that is, as packing
becomes tighter, the value of ae must decrease. Thus, ae
should not only depend on the headgroup structure but also
on interfacial hydration and surfactant concentration and
counterion type. Counterions added as salt are known to
induce the sphere-to-rod transition of micelles and increasing the counterion concentration must affect ae as well.
That is, the value of ae reflects the overall balance of
forces determining the aggregate size and shape at thermodynamic equilibrium. As pointed out by Nagaragan, ae is
a thermodynamic parameter.74 Romsted has demonstrated
that it also depends on interfacial hydration and on headgroupcounterion interactions in ionic aggregates.3

5.2.1 Contributions to the free energy of


micellization
The standard free energy of transfer of a unimer from
the aqueous region to the micelles, G (expressed as
G instead of chemical potential) can be divided into
three separate contributions as described by Tanford19 and
elegantly reviewed by Nagarajan74 :

G = GTransfer + GInterface + GHead

(8)

where the three terms on the right-hand side of (8) are the
free energies of transfer from bulk water to the micelle,
respectively, for the unimer tail to the micellar core,
for hydrocarbon/water contact in the interfacial region,
and for transfer of headgroups and counterions. Note
that an unknown and probably a variable fraction of
the surfactant tail is in contact with water and does not
contribute to G Transfer . This division of the overall free
energies of transfer is consistent with the equilibrium
micelle structure, Figure 5(b), and the overlapping radial
distribution functions of water and oil in Figure 5(c).
Similar expressions containing additional terms can be
written for other mesophases such as microemulsions.
The driving force, G Transfer
The primary contributor to G Transfer is the hydrophobic
effect that drives aggregation by minimizing the hydrocarbonwater contact. In the aqueous region, water molecules
surround the tails and the headgroups and counterions of
the surfactant unimers (Figure 5). Water molecules in the
immediate vicinity of the tails are believed to be oriented
such that they hydrogen bond with each other and not

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly

includes the interactions of surfactant methylenes within


the aggregate that remain in contact with water. G Head
includes the interactions of headgroups and a fraction of
bound counterions, , with each other and with water.
Changes in interactions between headgroups and counterions on going from dilute unimers in water to a high
local concentration of headgroups and counterions in the
interfacial region may also contribute to the entropy and
enthalpy of micellization, but the origin and size of these
contributions are currently unknown. The traditional explanation for the balance of forces controlling the micelle size
and shape equates the net charge of the interfacial region
produced by electrostatic repulsions between the charged
headgroups, the fractional charge = , with the balancing
force to the hydrophobic affect.14 What is not known is
how many coulombic interactions contribute to the total
balancing force or positive free energy terms compared to
hydration and ion-specific interactions between headgroups
and counterions.
The free counterions form an electrical double layer in
which the counterion concentrations around each micelle
decrease in a PoissonBoltzmann distribution into the
aqueous phase.16, 41 Figure 6 illustrates the double layer
and the radial distributions of counterions at different salt
concentrations obtained by solving the PoissonBoltzmann
equation. Note that the thickness of the double layer
depends on the ion salt concentration. The graph also illustrates a two-site model for ion distribution used in the
pseudophase models to describe measured ion distributions
in solutions of ionic association colloids, that is, counterions
are either bound or free (see below). However, explanations
based only on coulombic interactions between headgroups
and counterions fail to account for commonly observed
trends in ion-specific effects, for example, the Hofmeister

the methylenes in the tails.75 These water molecules are


assumed to be more ordered than nearby water molecules
in the bulk aqueous pseudophase.75 For single-chain surfactants, cooperativity is high, that is, 30100 unimers form
one spherical micelle at surfactant concentrations just above
the cmc. Aggregation of such large numbers of unimer tails
into the totality of the micelles in solution releases a significant, but unknown, amount of water into the surrounding
aqueous pseudophase producing a significant increase in
more disordered, bulk water.
Determinations of the thermodynamic parameters for
micellization in the vicinity of room temperature support
the above interpretation. The major contributor to the free
energy of micellization, G m per unimer, is from the
entropy of micellization, S m , Table 3.15, 76 Note that
the value of TS m is numerically similar to the value
of G m in most of the examples. H m is generally
small, probably because the dispersion interactions of
molecules of hydrocarbon with water and hydrocarbon
with hydrocarbon are similar. The added salt has a small
but real effect on H m , for example, DTAB, 0.1 M
NaBr, and increasing the temperature from 5 to 50 C
changes H m by 10 kJ mol1 . There are several reports
that show that, at higher temperatures, the H m becomes
more negative and T S m less positive.77 The reason for
the compensating temperature dependence of H m and
T S m is not understood.75

5.2.2 The balancing force, G Interface + G Head


The totality of the interactions contributing to G Interface +
G Head are as follows: the repulsive force that balance
the interactions contributing to G Transfer , the attractive
force, or the hydrophobic effect. The G Interface term
Table 3

13

Thermodynamic parameters of micellization in aqueous solution for some representative surfactants under various conditions.

Gm = Hm T Sm

Surfactant

Conditions: temperature/methoda

SDS
DTAB
DTAB
DTAB
DTAB
C12 H25 Pyr+ Br
C12 H25 Pyr+ Cl
Triton X 100
C12 E6
C12 H25 N+ (CH3 )2 O
Sodium taurocholate

25 C (cal)
25 C (cal)
0.1 M NaBr, 25 C (cal)
5 C (cal)
50 C (cal)
25 C (temp)
25 C (temp)
25 C (temp)
25 C (temp)
30 C (temp)
20 C (cal)

a Cal,

G m kJ mol1

H m kJ mol1

21.8
35.6
34.8
34.2
37.4
38.0
35.2
30.6
33.0
25.9
24.1

+0.4
1.4
2.1
+5.3
4.9
14
+2
+8.8
+16
+7
1.3

T S m kJ mol1

+22.2
+34.2
+32.7
+39.5
+32.5
+25
+38
+39.4
+49
+33
+22.8

calorimetry; temp, temperature change of cmc.


of a decimal point on a value of a parameter indicates additional significant figures are uncertain.

b Absence

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Reference
76
76
76
76
76
15
15
76
15
15
76

14

Concepts
Core

Surface
+

Aqueous phase

1.0 M

log[X ]

0.1 M
1

0.01 M
2

5.2.4 Ion-specific effects, the ion-pair/hydration


model

2
Micelle radii

Hofmeister, thousands of reports show that specific colloidal or biological properties often correlate with the
increasing radius or hydration-free energy of a series of
ions, for example, halide ions or alkali metal ions.1 These
two trends correlate with each other because, for ions of
the same valence, larger ions have lower free energies
of hydration.81 For example, the cmc values of surfactants depend on counterion type.15 For dodecyl sulfate
micelles with alkali metal counterions, the cmc values
decrease in the order: Li+ > Na+ > K+ > Rb+ > Cs+ ;
and for cationic surfactants with quaternary ammonium
headgroups, the cmcs values decrease in the order: F >
Cl > Br > I .15 Periodically, other orders are observed
as well when the headgroup type is changed in micelles and
other charged interfaces.82 In addition, ion-specific effects
on association colloid properties are often observed at elevated salt concentrations above 0.15 M, and the thickness
of the the electrical double layer approaches the thickness of the interfacial region and electrostatic headgroup
repulsions are completely screened by counterions and no
longer contribute to the forces balancing the hydrophobic
effect.2

Figure 6 Top: Cartoon illustrating locations of an organic substrate (see below), surfactant, coions, and counterions across a
small cross section of a cationic micelle interface. Bottom: Illustration of radial counterion distributions at three salt concentrations as described by the solution of the PoissonBoltzmann
equation in spherical symmetry (solid lines) and by the PIE model
assuming that = 0.75, cmc = 0, and interfacial counterion con a typical assumed
centration = 4 M (broken lines).  = 2.4 A,
width of the reaction region in the PBE model. (Reproduced from
Ref. 71 American Chemical Society, 1991.)

series, on aggregate properties, such as the cmc, aggregation number, Krafft temperature, sphere-to-rod transition,
counterion exchange, and rates of reactions speeded and
inhibited by ionic micelles.

5.2.3 Ion-specific effects


In 1888, Hofmeister showed that the solubility of a protein in water depends on the salt type and correlates
with the ion size and hydration. Both anions and cations
had independent orders. Small hydrophilic ions tend to
salted-out proteins, whereas larger, more hydrophobic less
strongly hydrated ions may salt-in the proteins.1, 7880 Since

Providing a coherent model for the Hofmeister series is


an unsolved problem in biological and colloid chemistry.2
A new approach, based on specific headgroupcounterion
interactions expressed as ion pairing, recognizes that ion
concentrations in interfacial regions of ionic association
colloids are high, 13 M and greater, as shown by product
yields from chemical trapping experiments.3 Note that in
Figure 5(c) the MD maximum interfacial of Br molarity
is about 1.3 M and about 2.5 M for the headgroups. At such
high ion concentrations, cations and anions can associate
reversibly in the interfacial region to form ion pairs.
Although they are neutral, but polar headgroupcounterion
pairs should have lower free energies of hydration than
free ions. Thus, interfacial ion-pair formation reduces the
interfacial water concentration resulting in an increase in
the interfacial packing permitting structural transitions, for
example, spherical to cylindrical geometry. Because ionpair formation is ion specific, it provides a straightforward
explanation for the dependence of the aggregate structure
on the counterion type, that is, less strongly hydrated
counterions form ion pairs at lower interfacial molarities.3

5.3

Balance of forces in nonionic aggregates

As discussed in Section 3.3.3, the hydrophobic effect


drives aggregation of nonionic surfactants but, unlike ionic

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly
micelles, it is balanced primarily by the free energy of
hydration interactions of the oxygens of the oligooxy
chains in the head group region with water contributing to the force balancing the hydrophobic effect.17 The
strengths of these interactions are characterized by the
number of water molecules per ethylene oxide unit in
the headgroup, the hydration number, which is about four
water molecules per ethylene oxide unit in micelles that
decreases with temperature as shown by chemical trapping
and self-diffusion83 and by five simulations.84 MD simulations show substantial hydrocarbonwater contact at the
core-interfacial region boundary that decreases with temperature, indicating that the oligooxy chains pack more closely
as the micelles grow.85 When the temperature exceeds the
cloud point, the hydration free energy is no longer sufficient to keep the surfactant in solution and it phase separates.17
The tendency of the oligooxy groups to dehydrate with
increasing temperature leads to an unusual phenomenon
with nonionic emulsions (very large, surfactant-coated
droplets that scatter visible light), the phase inversion
temperature (PIT).56 Warming and mixing a two-phase
mixture of an oil-in-water emulsion with an excess oil
phase on top (oil is less dense than water) leads to gradual
incorporation of the oil in the emulsion as the temperature
increases until a single emulsion is observed at the PIT,
whose value correlates with the HLB of the nonionic
surfactant. Above this temperature, a water layer appears
at the bottom and a water-in-oil emulsion phase at the
top. This transition requires a shift in the balance of forces
controlling the droplet structure such that the droplets invert
from oligooxy chains on the outside of the droplets (oilin-water emulsion) to the inside of the droplets (water-inoil emulsion) above the PIT. This reversal of curvature
requires tighter packing of the oligooxy chains, which
occurs because the oligooxy chains lose water of hydration.
Similarly, addition of large quantities of salt, such as NaCl,
on the order of 1 M and higher, also induces this inversion
of the droplet structure.62

6.1

ASSOCIATION COLLOID EFFECTS


ON CHEMICAL REACTIVITY,
AN INTRODUTION

15

distributions and rate constants.3, 63, 71, 8691 The specific


aims in studying chemical reactions in association colloids are multiple, for example: to develop surfactants
as catalysts,89 to mimic and therefore better understand
enzymatic catalysis especially the concentration effect,92, 93
to probe the physical and chemical properties of aggregates,3, 63 to decontaminate chemical warfare agents,94 and
to develop association colloids as nanoreactors for synthesis and green chemistry.95, 96 The complete development
of these aims requires the understanding of the factors that
contribute to catalytic activity in association colloids. Many
reactions have been studied using a wide range of surfactants under a variety of conditions, that is, aggregate
structures (micelles, microemulsions, and vesicles), temperature and pH ranges, salt concentrations and counterion
type, and in the presence of additives such as alcohols.
The best starting points are reviews,63, 71, 8688, 90, 91 and a
book.89
Pseudophase models work for several reasons: (i) Reactions in association colloids can be carried out under
conditions of dynamic equilibrium. Thus the totality of
the interfacial regions of all the aggregates in micelle,
microemulsion, or vesicle solutions, and the totalities of
their oil and water regions can be modeled as single interfacial, oil, and water reaction volumes of uniform properties with a separate rate constant for the reaction in each
volume, Scheme 4. (ii) The requirement of dynamic equilibrium is met because the rate constants for diffusion of
ions and molecules in association colloid solutions are near
the diffusion-controlled limit. For example, the entrance and
exit rate constants in micellar solutions in Table 1 are orders
of magnitude faster than the example rate constants for thermal bimolecular reactions in micellar solutions in Table 4.
Many additional examples are compiled in reviews.63, 86
(iii) Measured rate constants for spontaneous reactions and

Aqueous
W

AW

NW

k2W

Product(s)

Interface
l

AM

NM

k2M

Product(s)

Oil
O

AO

NO

k2O

Product(s)

Pseudophase kinetic models

Modeling of association colloid effects on the rates of


chemical reactions is based on the pseudophase model.
The results from many studies illustrate the explanatory
power of this approach and how chemical reactivity in
association colloids is interpreted in terms of component

Scheme 4 A cartoon of a small section of a self-assembled


surfactant aggregate in an association colloid or emulsion showing
the aqueous, interfacial, and oil regions. Substrate A and reactant
N are in dynamic equilibrium between all three regions. A reaction
between A and N may occur in one, two, or all three regions
depending on their solubilities in the regions and their volume
fractions, .

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

16

Concepts

Table 4 Selected examples of parameters used to fit kobs versus [ST ] profiles of bimolecular reactions in micellar solutions.a Values
of , VM , and temperature are not included.
Substrate

Surfactant

KA M1

KN X

k2 M /k2 W

Br
Cl
OH

CTAOMsb
CTAOMs
CTPOHc

65
65
65

2.5
2

1.7
0.41
0.24

OH
OH
OH

CTABrd
CTPABre
CTBABrf

1600
1600
1600

14
14
14

2.7
3.6
4.7

OH
OH
OH

CTANO3 g
CTABrd
CTAOH

285
295
200

37
32

1.9
2.5
1.1

BH4 f
BH4

CTABr
CTACld

70
70

2
4

0.028
0.036

Nucleophilic substitution
SO3CH3

Cl
NO2

NO2

Eliminations by OH
CH2CH2Br

Borohydride reduction
O
H3CO

Ref. 63.
Cetyltrimethylammonium mesylate.
c Cetyltri-N-propylammonium hydroxide.
d
Cetyltrimethylammonium X, X = OMs, OH, NO3 , Br , Cl .
e Cetyltri-N-propylammonium hydroxide.
f
Cetyltri-N-butylammonium bromide.
g Borohydride.
b

calculated bimolecular rate constants are essentially constant aqueous solutions of the surfactant and salt (up to
about 0.2 M of either). This constancy indicates that the
properties of micelles as reaction media are not very sensitive to solution composition. (iv) Association constants of
neutral substrates or ion exchange constants for ionic reactants can be used to relate the distribution of a reactant to its
stoichiometric concentrations in solution. (v) Substrate concentrations can be kept small, 104 to 105 M, and that
of the second reactant in about 10-fold excess; such chemical reactions can be run under first-order conditions and
neither reactant perturbs the properties of the association
colloid significantly.
The next sections introduce the application of pseudophase kinetic models to spontaneous and second-order
reactions in micelles as a general approach for treating
chemical reactivity in micelles.

6.2

Spontaneous or first-order reactions

In a bulk homogeneous solution, the value rate constant of


a spontaneous or first-order reaction, kobs , depends only on

the properties of the reaction medium, for example, benzene


is a less polar solvent than water. Organic substrates, either
neutral or ionic, are polar and there is no evidence of
the reaction occurring in the nonpolar micellar core, and
the micellar rate constant is generally assumed to reflect the
polarity of the interfacial region. In addition, because the
volume of a micelle increases with the cube of its radius,
the interfacial region is about 50% of the total micellar
volume. However, the oil region may be substantial in
microemulsions because oil is added to the solution and
some fraction of the uncharged substrate may be located in
the oil region.
Aqueous micellar solutions contain two separate reaction
media, water and micelles, and because the components are
in dynamic equilibrium, the value of kobs can be expressed
as the sum of the rate in each region or pseudophase, (9).
Observed rate = kobs [AT ] = kW [AW ] + kM [AM ]

(9)

Above the cmc, the observed rate and kobs at any one surfactant concentration is determined by the medium properties
of the micelles, kM ,91 and water, kW , and the concentration
of the substrate A in each region. Equation (10) is a typical association constant used to characterize the distribution

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly
of a substrate, A, between micelles and water. In general,
A is larger for less water-soluble substrates. Equations (9
and 10) are combined with mass balance equations for the
surfactant, (2), and the substrate, [AT ] = [AW ] + [AM ], and
solved for kobs in terms of [ST ], (11). Note that kobs only
depends on the aqueous and micellar rate constants, KA and
[ST ], and does not depend on the concentration of substrate.
The rate constants are in units of s1 and KA is in units of
M1 .
[AM ]
KA =
Aw + SM  AM
[AW ][SM ]
kW + kM KA ([ST ] cmc)
kobs =
(1 + KA ([ST ] cmc)
KA

(10)
(11)

For a reaction that is speeded by micelles, (11) describes


a sigmoidal curve that rises at the cmc and reaches a
plateau when A is completely micellar bound, that is, when
KA ([ST ]cmc)> 1 and when kM KA ([ST ]cmc)> kW such
that kobs = kM . Below the cmc, kobs = kW . The value of KA ,
is obtained by fitting the kobs versus [ST ] data. Traditional
methods, for example, UVvisible, cannot be used to
estimate KA using (10) because the substrate is reacting
while the measurements are in progress.
The change in kobs may be small, for example, a factor
of 2, or dramatic, 103 or greater, and it generally correlates
with the effect of solvent polarity on the reaction. No
examples are reported of substrates reacting in nonpolar
micellar core and the value of kM is generally assumed to
reflect the polarity of the interfacial region.

6.3

Second-order reactions and the pseudophase


ion exchange (PIE) model

The pseudophase kinetic models for speeded or inhibited


bimolecular, second-order, reactions are more complex.
Here the focus is on reaction between a neutral organic
substrate and a reactive counterion in micellar solutions in
the absence of oil (0 = 0, Scheme 4). Micellar effects
on reactions of substrates with reactive counterions are
important because they illustrate the general differences
of micellar effects on spontaneous and bimolecular reactions and also how specific counterion effects influence
the results. Pseudophase models also work for bimolecular reactions between two uncharged organic substrates
and third-order reactions,63, 91, 97 reactions in vesicles and
microemulsions, which may include partitioning into and
reaction in the oil region,98 reactions of substrates with
an ionizable (e.g., deprotonatable) second reactant, and the
effect of association colloids on indicator equilibria.86
The observed or measured rate constant, kobs , under
pseudo first-order conditions of a bimolecular reaction

17

between a substrate, A, and a second reactant, N, in


micelles, Figure 3 (lower left side, no added oil), Figure 4
(micelle mic region, room temperature) is the sum of
the rates in the micellar and aqueous regions:
Observed rate = kobs [AT ] = k2 [AT ][NT ]
= k2 W [AW ][NW ] + k2 M AM NM

(12)

where subscripts T, W, and M represent the stoichiometric,


aqueous, and interfacial concentrations. The second-order
rate constants, k2 , k2 W , and k2 M , in units of M1 s1 , are for
the observed, aqueous, and micellar reactions, respectively.
In general, reactions are run under first-order conditions
with [NT ] > [AT ] such that the concentration [NT ] remains
constant throughout the reaction, and kobs = k2 [NT ], (12).
The value of k2 W is set equal to the rate constant in water
in the absence of micelles. The value of k2 M , which cannot
be estimated independently, is estimated by fitting kobs
versus [ST ] profiles using independent estimates of the other
parameters.
The concentration units of A and N in the aqueous and
micellar pseudophases require special consideration. At low
micelle concentrations, for example, <35% of the total
solution volume, a common experimental condition, the
molarities of A and N in the aqueous pseudophase can
be expressed in stoichiometric units because the micelles
occupy only a small fraction of the total pseudophase
volume, (12), square brackets. However, the rate of the
reaction in the micellar pseudophase depends on their
concentrations in the interfacial region and not on the whole
solution volume. This point is crucial. The concentrations of
A and N in the interfacial region are expressed in moles/liter
of interfacial volume. Both A and N are assumed to be
located in the interfacial region because they are polar or
ionic. Equation (13) defines the relationship between the
interfacial concentration of N, NM , and the stoichiometric
concentrations of N and S in micelles:
NM =

[NM ]
[SM ]VM

(13)

where VM is the molar volume of reaction and is generally


set equal to the estimated volume of the interfacial region
in liter/mole. The value of NM defines how concentrated
reactant N is in the interfacial compared to its stoichiometric
concentration, [NW ] in the absence of added N as salt (see
below). An equation of the same form is used to describe
the interfacial concentration of A. The value of Vm is
typically about 0.15 and can be estimated by assuming that
micelles have a density of 1, and that the interfacial region
occupies about 50% of the total micellar volume.
The distribution of A (and any uncharged substrate
or reactant) in bimolecular reactions is defined in (11).

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

18

Concepts

The distribution of a reactive counterion is described by


ion exchange constant because substantial experimental
work has demonstrated that micellar and other aqueous
ionic interfaces act as selective ion exchangers.88, 99103
For example, reactive counterions, for example, H+ in
anionic micelles exchange with other cations such as Li+
or Cs+ and reactive anions, for example, OH or CN in
cationic micelles are displaced by nonreactive counterions.
Their distributions are generally described by ion exchange
constants, (14):

200
a

krel

150

N
A

N
A
X
N
X

a
b
c
d

KA M1
10000
1000
100
10
X

100

N
A

KN X

NM + XW  NW + XM

KN X

[NW ]XM
=
NM [XW ]

k2 =

k2 M KA /VM ([ST ] cmc)


(KA ([ST ] cmc) + 1)([NT ] + [XT ]KN X )
k2 w
(15)
+
(KA ([ST ] cmc) + 1)

The two terms, one for the reaction in micelles (right


hand side, first term) and water (right-hand side, second
term of this equation are primarily composed of independently measurable or estimable constants, KA , , VM , cmc,
k2 W , KA , and KN X and experimentally fixed concentrations, [ST ], [NT ], and [XT ]). Above the cmc, the second
term rapidly diminishes, which is important for catalyzed
reactions (most studies), but makes a significant contribution for inhibited reactions (not considered). The first term
defines the relationships between k2 and [ST ] and [XT ].[XT ]
= [ST ] because X is added as a counterion with surfactant and [NT ] is often small in comparison. The first term
describes a maximum in k2 with increasing [ST ] because
the numerator is first order and the denominator is second
order in surfactant concentration. Figure 7 shows a simulated kinetic plot based on (15) (real data is seldom this
uniform). It illustrates the effect of the added cationic surfactant on a reaction between the neutral substrate, A, an
anionic reactant, N, and the anionic nonreactive counterion,
X, with increasing [ST ] and the effect of increasing KA on

N
X

(14)

Measured values for KN X are numerically similar to those


for loosely cross-linked ion exchange resins104 and to the
examples in Table 4. The order of effectiveness of X in
displacing N from the micellar interface generally follows
a Hofmeister series for both cations and anions, that is, KN X
increases as the size of the X increases (and is more weakly
hydrated) relative to N. Combining the rate expression with
the equations for the association of A to micelles and ion
exchange of N and X, the degree of counterion binding, ,
for X, the definition of NM , and an equivalent definition
for XM , and the needed mass balance equations gives (after
some struggle) (13)86 for the observed second-order rate
constant, k2 :

A
X

N
A

N
A

50
X

N
X

d
0

0.01

0.02
0.03
[ST] M

0.04

0.05

Figure 7 Effect of increasing substrate hydrophobicity on krel =


k2 /k2 W , which compares the increase in k2 produced by concentrating reactants in micelles to the rate constant in water. krel = 1
at the cmc and below. Other parameter values are as follows:
cmc = 0.002 M, [NT ] = 0.01 M, KN X = 4, k2 M = k2 W , = 0.7,
and Vm = 0.15 M1 . KA values are listed in the figure. The cartoon beakers illustrate the transferring and concentration of S and
N in micelles above the cmc and their dilution in the micellar
pseudophase at high [ST ]. (Redrawn from Ref. 71. American
Chemical Society, 1991.)

krel , the relative rate enhancement ratio of the observed


second-order rate constant to the second-order rate constant in water, krel = k2 /k2 W versus [ST ], assuming that
k2 W = k2 M .
Many examples of such plots are listed in the references cited in the first paragraph of Section 6.1. The rate
maximum in the krel versus [ST ] M profile is a crossover
point between two trends that both depend on increasing
[ST ]. They are illustrated by the cartoon beakers containing
unimer, micelles, substrate, anionic reactant, and nonreactive counterion. Below the cmc, no micelles are present
and k2 = k2 W . Above the cmc, increasing [ST ] increases
the micelle concentration and the fraction of A and N in
the interfacial region of the micelles. The rate of increase
in their bound fractions depends on the values of KA for
A and KN X for N and X. The simulations in Figure 7
show that the larger the KA , that is, for more hydrophobic substrates with the same functional group, the more
rapidly krel increases with increasing [ST ], for example,
KA = 10000. Concurrent with the increase in the fraction
bound of A and N, their molarities in the interfacial region
in moles per liter of interfacial volume are decreasing continuously, as illustrated by the last beaker that contains
more micelles, but the same stoichiometric concentrations
of A and N as the beaker near maximum in k2 . Thus, the

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly
rate maximum is a crossover point where the value of krel
changes from being determined by transfer of reactants
from bulk solution to the smaller volume of the micellar
pseudophase to dilution of the reactants within the micellar
pseudophase.
At very high [ST ] M, when A is completely micellar
bound, (15) simplifies and all four curves in Figure 7 tend
toward a single curve and k2 equals:
k2 =

K2M /VM
([NT ] + [XT ]KNX )

variety of counterions. The trends in their values generally


follow a Hofmeister series and are direct evidence for the
importance of ion-specific effects on micellar properties.
Table 4 shows a few examples of reactions studied for
which most of the parameters are available for bimolecular
reactions studied in micelles. As noted above, all the
constants in (15) can be determined independently, except
the interfacial rate constant, k2 M , which must be obtained
from the simulation of the kinetic data.

(16)

6.4

The value of k2 always decreases with increasing [ST ],


when [ST ] = [XT ] in (16). A value for k2 M can be estimated
from (16) by using experimental values for k2 , [NT ], [XT ],
, and KN X and an estimated VM or by fitting the entire k2
versus [ST ] profile with (15).
Figure 8 shows simulations of the effect of added nonreactive counterions on krel at constant [ST ] M based
on (15). Adding X decreases N in the micellar interface
by ion exchange, which decreases krel by reducing [NM ],
as illustrated by the cartoon beakers. At [ST ] + [XT ] = 0.2,
krel decreases 20-fold (curve a) when KN X = 1. When
KN X = 24, krel decreases over a 100-fold (curve d). Many
measured ion exchange constants have been obtained for a

Reactive counterion surfactants

Reactive counterion surfactants have only one counterion and it is also the reactant, for example, OH in
cetyltrimethylammonium hydroxide (CTAOH) micelles.
The stoichiometric concentration of micellized N, [NM ]
increases incrementally with added surfactant, but the interfacial molarity, NM , is constant because, in the original
development of pseudophase models, the assumption was
made that [NM ]/[SM ] = and that is constant and
independent of surfactant concentration and the concentration of any added salt. Equation (15) reduces to (17).
Equation (17) has the same mathematical form as (11) for
first-order reactions.
k2 =

80

A
N
A
N
A
N

60

a
b

K NX
1
4

c
d

12
24

krel

40

X N
A
N
X

A
X
A
X

N
X

c
20

k2 W [NW ] + k2 M KA /VM ([ST ] cmc)


KA ([ST ] cmc)

0.1
([ST] + [XT]) M

0.2

Figure 8 Specific counterion effects on krel as a function of


added X for typical KN X values (see figure) at surfactant concentration = 0.02 M and KA = 1000 M1 . Other parameters as
in Figure 7. The surfactant and added salt have the same counterion in curve b. The cartoon beakers illustrate the displacement
of N by added X with increasing [XT ]. (Redrawn from Ref. 71.
American Chemical Society, 1991.)

(17)

This make sense because, if the interfacial reactive counterion concentration is constant, (13), then it was reasonable
to assume that the ratio of bound counterions to micellized surfactant, , and VM are constant. This prediction
has been observed many times when kobs increases with
added [CTANT ] until all the substrate, A, is micellar bound
and then kobs becomes constant.91, 105 However, this prediction has also failed when a significant amount of counterion
is added as salt, above 0.2 M to more than 1 M.106108
The experimental results at high concentrations of added
reactive counterion are qualitatively consistent with the
assumption that the total interfacial counterion concentration is given by
NM =

d
0

19

[NM ]
+ [NW ]
[SM ]VM

(18)

This equation states that the interfacial molarity of counterion is equal to the initial molarity in the absence of added
salt, first term, plus the molarity of the stoichiometric concentration of the counterion added as salt, second term. This
interpretation qualitatively fits both kinetic and chemical
trapping results above about 0.2 M of added salt and indicates that the micellar interfaces (and by implication other
association colloid interfaces) do not saturate with counterions as originally assumed in pseudophase kinetic models.3

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

20

Concepts

CONCLUSIONS

Current understanding of the self-assembly of surfactants


in solution has come a long way since McBain was chided
a century ago for suggesting that soap molecules spontaneously aggregate in water. There is now overwhelming
evidence that nonionic, zwitterionic, and ionic surfactants
(and amphiphiles, soaps, and detergents) self-assemble to
form spherical aggregates in dilute aqueous solution with
the surfactant tails in the core, and the headgroups arrayed
over the surface in contact with water. Increasing the surfactant concentration produces a variety of structures in other
mesophases, for example, rods, hexagonal, cubic, lamellar,
bicontinous, vesicular, and reverse structures. Much has
been learned about how to characterize micellar properties, for example, cmc values, ionization degrees, aggregation numbers, and the sizes and shapes of the aggregates. The pseudophase model has proved to have broad
general utility in interpreting association colloid properties
including their effects on the rates of chemical reactions.
The current understanding of the balance of forces controlling the aggregate size and shape must be modified to
include specific interactions between surfactant headgroups
and counterions and water, in the interfacial regions of
the aggregates. Aggregation is driven by the hydrophobic
effect and balanced by interactions between headgroups,
counterions, and water in the interfacial region. Recent evidence is consistent with the formation of ion pairs between
headgroups and counterions in the interfacial regions of
micelles with the concomitant release in water, an increase
in interfacial packing, and structural transitions of micellar aggregates. Progress will come from careful probing
of the compositions of aggregate interfaces over a wide
range of solution compositions via chemistry and continued
application of molecular dynamics simulations and quantum
mechanics.

ACKNOWLEDGMENTS

2. B. W. Ninham and P. Lo Nostro, Molecular Forces and


Self Assembly in Colloid, Nano Sciences and Biology,
Cambridge University Press, Cambridge, 2010.
3. L. S. Romsted, Langmuir, 2007, 23, 414.
4. http://en.wikipedia.org/wiki/soap, Wikipedia, May 18, 2011.
5. E. Carretti, R. Giorgi, D. Berti, and P. Baglioni, Langmuir,
2007, 23, 6396.
6. J. W. McBain and H. E. Martin, J. Chem. Soc., Trans. 1 ,
1914, 105, 957.
7. http://en.wikipedia.org/wiki/Surfactant, Wikipedia, May 18,
2011.
8. F. M. Menger, Acc. Chem. Res., 1979, 12, 111.
9. http://en.wikipedia.org/wiki/Amphiphile, Wikipedia, May
18, 2011.
10. F. M. Menger, Langmuir, 2011, 27, 5176.
11. P. Somasundaran, Encyclopedia of Surface and Colloid
Science, 2nd ed., ed. P. Somasundaran, CRC Press, Boca
Raton, 2006.
12. Kirk-Othmer Encyclopedia of Chemical Technology, 5th ed.,
ed. A. Seidel, Wiley-Interscience, Hoboken, 2007.
13. Surfactant Science Series, http://www.crcpress.com/ ecommerce product/book series.jsf?series id=1766, CRC Press,
Boca Raton, 2011.
14. J. Israelachvili, Intermolecular and Surface Forces, 2nd ed.,
Academic Press, London, 1991.
15. M. J. Rosen, Surfactants and Interfacial Phenomena, 3rd
ed., John Wiley & Sons, New York, 2004.
16. D. F. Evans and H. Wennerstrom, The Colloidal Domain:
Where Physics, Chemistry, Biology and Technology Meet,
VCH Publishers, New York, 1994.
17. B. Jonsson, B. Lindman, K. Holmberg, and B. Kronberg,
Surfactants and Polymers in Aqueous Solution, John Wiley
& Sons, Chichester, 1998.
18. A. W. Adamson and A. P. Gast, Physical Chemistry of
Surfaces, 6th ed., J. Wiley & Sons, New York, 1997.
19. C. Tanford, The Hydrophobic Effect: Formation of Micelles
and Biological Membranes, 2nd ed., Wiley, New York,
1980.
20. T. A. Camesano and R. Nagarajan, Colloids Surf., A, 2000,
167, 165.
21. E. Ruckenstein, Adv. Colloid Interface Sci., 1999, 79, 59.

Financial support comes from the NSF Organic and Macromolecular Chemistry Program (CHE-0 840916) and the
United States Department of Agriculture (CSREES AFRI
93430). Many collaborators have made this work possible,
especially C. A. Bunton. Eric Romsted patiently drew a
number of Figures and Schemes and without Jean Romsted,
well, the chapter would not exist.

REFERENCES
1. K. D. Collins and M. W. Washabaugh, Q. Rev. Biophys.,
1985, 18, 323.

22. M. Manciu and E. Ruckenstein, Adv. Colloid Interface Sci.,


2004, 112, 109.
23. B. Stephenson, A. Goldsipe, K. Beers, and D. Blankschtein, J. Phys. Chem. B , 2007, 111, 1045.
24. B. Stephenson, A. Goldsipe, K. Beers, and D. Blankschtein, J. Phys. Chem. B , 2007, 111, 1025.
25. B. Stephenson, A. Goldsipe, K. Beers, and D. Blankschtein, J. Phys. Chem. B , 2007, 111, 1063.
26. J. N. Israelachvili, D. J. Mitchell, and B. W. Ninham, J.
Chem. Soc., Faraday Trans. 2 , 1976, 72, 1525.
27. J. A. Pelesko, Self-Assembly: The Science of Things that Put
Themselves Together, Chapman & Hall/CRC, Boca Raton,
2007.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly

21

28. J. W. Steed and J. L. Atwood, Supramolecular Chemistry,


2nd ed., John Wiley & Sons, Chichester, 2009.

54. M. Fanun, eds., Microemulsions: Properties and Applications, CRC Press, Boca Raton, 2009.

29. R. Dias and B. Lindman, DNA Interactions with Polymers


and Surfactants, John Wiley and Sons, Hoboken, 2008.

55. B. Jonsson, P.-G. Nilsson, B. Lindman, et al., Principles of


phase equlibria in surfactant-water systems, in Surfactants
in Solution, Vol. 1, eds. K. L. Mittal and B. Lindman,
Plenum, New York, 1984, p. 3.

30. M. Lazzari, G. Liu, and S. Lecommondaux, eds., Block


Copolymers in Nanoscience, Wiley-VCH, Weinheim, 2006.
31. F. Vogtle, G. Richardt, and N. Werner, Dendrimer Chemistry: Concepts, Syntheses, Properties, Applications, WileyVCH, Weinheim, 2009.
32. T. Imae, ed., Advanced Chemistry of Monolayers at Interfaces: Trends in Methodology and Technology, Academic
Press, London, 2007.
33. K. S. Birdi, Self-Assembly Monolayer Structures of Lipids
and Macromolecules at Interfaces, Kluwer Academic, New
York, 1999.
34. F. M. Menger, Comptes Rendus Chimie, 2009, 12, 54.
35. F. M. Menger and J. S. Keiper, Angew. Chem., Int. Ed.
Engl., 2000, 39, 1906.
36. R. Zana and J. Xia, Gemini Surfactants: Synthesis, Interfacial and Solution-Phase Behavior, and Applications, Vol.
117, Marcel Dekker, New York, 2004.

56. K. Shinoda and S. Friberg, Emulsions and Solubilization,


John Wiley & Sons, New York, 1986.
57. D. Balzer and H. Luders, eds., Nonionic Surfactants: Alkyl
Polyglucosides, Marcel Dekker, Boca Raton, 2000.
58. M. J. Schick, Nonionic surfactants: physical chemistry, in
Surfactant Science Series, Vol. 23, eds. M. J. Schick and
F. M. Fowkes, Marcel Dekker, New York, 1987.
59. D. J. Mitchell, G. J. T. Tiddy, L. Waring, et al., J. Chem.
Soc., Faraday Trans. 1 , 1983, 79, 975.
60. M. Jonstromer, B. Jonsson, and B. Lindman, J. Phys.
Chem., 1991, 95, 3293.
61. L. S. Romsted and J. Yao, Langmuir, 1999, 15, 326.
62. J. Yao and L. S. Romsted, Langmuir, 2000, 16, 8771.

37. J. P. Priebe, M. L. Satnami, D. W. Tondo, et al., J. Phys.


Chem. B , 2008, 112, 14373.

63. G. Savelli, R. Germani, and L. Brinchi, Reactivity control


by aqueous amphiphilic self-assembling systems, in Reactions and Synthesis in Surfactant Systems, Vol. 100, ed.
J. Texter, Marcel Dekker, New York, 2001, p. 175.

38. G. Cevc, ed., Phospholipid Handbook , Marcel Dekker,


New York, 1993.

64. E. A. G. Aniansson, S. N. Wall, M. Almgren, et al., J.


Phys. Chem., 1976, 80, 905.

39. D. Madenci and S. U. Egelhaff, Curr. Opin. Colloid Interface Sci., 2010, 15, 109.

65. R. Zana, Dynamics of Surfactant Self-Assembly: Micelles,


Microemulsions, Vesicles, and Lyotropic Phases, Vol. 125,
CRC Press, Boca Raton, 2005.

40. W. H. Binder, V. Barragan, and F. M. Menger, Angew.


Chem., Int. Ed. Engl., 2003, 42, 5802.

66. L. S. Romsted, Ph.D. thesis, Indiana University, 1975.

41. P. C. Hiemenz and R. Rajagopalan, eds., Principles of Colloid and Surface Chemistry, CRC Press, Boca Raton, 1997.

67. I. M. Cuccovia, I. N. da Silva, H. Chaimovich,


L. S. Romsted, Langmuir, 1997, 13, 647.

42. B. Gohain and R. K. Dutta, J. Colloid Interface Sci., 2008,


323, 395.

68. L. Sepulveda and J. Cortes, J. Phys. Chem., 1985, 89, 5322.

and

43. D. Balzer, Langmuir, 1993, 9, 3375.

69. C. Gamboa, H. Rios, and L. Sepulveda, J. Phys. Chem.,


1989, 93, 5540.

44. G. G. Warr, T. N. Zemb, and M. Drifford, J. Phys. Chem.,


1990, 94, 3086.

70. S. Manet, Y. Karpichev, D. Bassani, et al., Langmuir,


2010, 26, 10645.

45. R. Zana, M. Benrraou, and B. L. Bales, J. Phys. Chem. B ,


2004, 108, 18195.

71. C. A. Bunton, F. Nome, F. H. Quina, and L. S. Romsted,


Acc. Chem. Res., 1991, 24, 357.

46. R. Zana, Adv. Colloid Interface Sci., 1995, 57, 1.


47. M. Abe and J. F. Scamehorn, eds., Mixed Surfactant Systems, Marcel Dekker, New York, 2005.
48. P. M. Holland and D. N. Rubingh, Mixed surfactant systems: An overview, in Mixed Surfactant Systems, Vol. 501,
eds. P. M. Holland and D. N. Rubingh, American Chemical
Society, Washington, DC, 1992, p. 2.

72. M. Jorge, Langmuir, 2008, 24, 5714.


73. C. D. Bruce,
M. L. Berkowitz,
L. Perera,
and
M. D. E. Forbes, J. Phys. Chem. B , 2002, 106, 3788.
74. R. Nagarajan, Langmuir, 2002, 18, 31.
75. N. T. Southall, K. A. Dill, and A. D. J. Haymet, J. Phys.
Chem. B , 2002, 106, 521.

49. S. B. Lioi, X. Wang, M. R. Islam, et al., Phys. Chem.


Chem. Phys., 2009, 11, 9315.

76. M. N. Jones and D. Chapman, Micelles, Monolayers, and


Biomembranes, Wiley-Liss, New York, 1995.

50. R. G. Laughlin, The Aqueous Phase Behavior of Surfactants, Academic Press, London, 1994.

77. D. F. Evans, Langmuir, 1988, 4, 3.

51. R. G. Larson, J. Chem. Phys., 1989, 91, 2479.

78. F. Hofmeister, Naunyn-Schmiedebergs Arch. Exp. Pathol.


Parmakol. (Leipzig), 1888, 24, 247.

52. R. Zana, ed., Surfactant Solutions: New Methods of Investigation, Marcel Dekker, New York, 1985.

79. W. Kunz, P. Lo Nostro, and B. W. Ninham, Curr. Opin.


Colloid Interface Sci., 2004, 9, 1.

53. R. Zana and E. W. Kaler, eds., Giant Micelles: Properties


and Applications, CRC Press, Boca Raton, 2007.

80. Y. Marcus, Chem. Rev. (Washington, DC, U. S.), 2009, 109,


1346.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

22

Concepts

81. Y. Marcus, Ion Solvation, John Wiley & Sons, Chichester,


UK, 1985.

105. L. Brinchi, P. Di Profio, R. Germani, et al., J. Colloid


Interface Sci., 2001, 236, 85.

82. Z.-M. He, P. J. OConnor, L. S. Romsted, and D. Zanette,


J. Phys. Chem., 1989, 93, 4219.

106. C. A. Bunton and L. S. Romsted, Reactive counterion surfactants, in Solution Behavior of Surfactants: Theoretical
and Applied Aspects, Vol. 2, ed. K. L. Mittal, Plenum Press,
New York, 1982, p. 975.

83. L. S. Romsted and J. Yao, Langmuir, 1996, 12, 2425.


84. G. Briganti, G. DArrugo, M. Marccarini, et al., Colloids
Surf., A, 2005, 261, 93.
85. F. Sterpone, C. Pierleoni, G. Briganti, and M. Marchi,
Langmuir, 2004, 20, 4311.
86. L. S. Romsted, Micellar effects on reaction rates and equilibria, in Surfactants in Solution, Vol. 2, eds. K. L. Mittal
and B. Lindman, Plenum Press, New York, 1984, p. 1015.
87. C. A. Bunton, J. Yao, and L. S. Romsted, Curr. Opin. Colloid Interface Sci., 1997, 2, 622.
88. C. A. Bunton and L. S. Romsted, Organic reactivity in
microemulsions, in Handbook of Microemulsion Science
and Technology, eds. P. Kumar and K. L. Mittal, Marcel
Dekker, New York, 1999, p. 457.
89. M. N. Khan, Micellar Catalysis, Vol. 133, CRC Press, Boca
Raton, 2006.
90. C. A. Bunton, Micellar rate effects upon organic reactions,
in Kinetics and Catalysis in Microheterogeneous Systems:
Surfactants in Science Series, Vol. 38, eds. M. Gratzel and
K. Kalyanasundaram, Marcel Dekker, New York, 1991,
p. 13.

107. M. F. S. Neves, D. Zanette, F. Quina, et al., J. Phys.


Chem., 1989, 93, 1502.
108. L. Brinchi, P. Di Profio, R. Germani, et al., Langmuir,
1997, 13, 4583.

FURTHER READING
Introductory books
M. J. Rosen, Surfactants and Interfacial Phenomena, 3rd ed.,
John Wiley & Sons, Inc., New York, 2004.
B. Jonsson, B. Lindman, K. Holmberg, and B. Kronberg, Surfactants and Polymers in Aqueous Solution, John Wiley & Sons,
Inc., Chichester, 1998.

91. C. A. Bunton, Adv. Colloid Interface Sci., 2006, 123126,


333.

Advanced/textbook/monograph

92. X. Zhang and K. N. Houck, Acc. Chem. Res., 2005, 38, 379.

C. Tanford, The Hydrophobic Effect: Formation of Micelles and


Biological Membranes, 2nd ed., Wiley, Inc., New York, 1980.

93. M. K. Jain, M. H. Gelb, J. Rogers, and O. G. Berg, Methods Enzymol., 1995, 249, 567.
94. R. A. Moss, A. T. Kotchevar, B. D. Park, and P. Scrimin,
Langmuir, 1996, 12, 2200.
95. R. Schoemacker and K. Holmberg, Reactions in organized
surfactant systems, in Microemulsions: Background, New
Concepts, Applications, Perspectives, ed. S. Stubenrauch,
Blackwell Publishing, Oxford, 2009, p. 148.
96. T. Dwars, E. Paetzold, and G. Oehme, Angew. Chem., Int.
Ed. Engl., 2005, 44, 7174.
97. C. A. Bunton, L. S. Romsted, and H. J. Smith, J. Org.
Chem., 1978, 43, 4299.

A. W. Adamson and A. P. Gast, Physical Chemistry of Surfaces,


6th ed., J. Wiley & Sons, Inc., New York, 1997.
J. Israelachvili, Intermolecular and Surface Forces, 2nd ed.,
Academic Press, London, 1991.
P. L. Yeagle, ed., The Structure of Biological Membranes, 2nd
ed., CRC Press, Boca Raton, 2004.
D. F. Evans and H. Wennerstrom, The Colloidal Domain: Where
Physics, Chemistry, Biology and Technology Meet, VCH Publishers, New York, 1994.

98. R. da Rocha Pereira, D. Zanette, and F. Nome, J. Phys.


Chem., 1990, 94, 356.

B. W. Ninham and P. Lo Nostro, Molecular Forces and SelfAssembly in Colloid, Nano Sciences and Biology, Encyclopedias/Book Series (example titles), Cambridge, Cambridge
University Press, 2010.

99. J. D. Morgan, D. H. Napper, and G. G. Warr, J. Phys.


Chem., 1995, 99, 9458.

P. Somasundaran, ed., Encyclopedia of Surface and Colloid Science, 2nd ed., CRC Press, Boca Raton, 2006.

100. F. H. Quina and H. Chaimovich, J. Phys. Chem., 1979, 83,


1844.

M. Fanun, ed., Microemulsions: Properties and Applications,


CRC Press, Boca Raton, 2008.

101. H. Chaimovich,
J. B. S. Bonilha,
M. J. Politi,
F. H. Quina, J. Phys. Chem., 1979, 83, 1851.

and

C. C. Ruiz, ed., Sugar-Based Surfactants: Fundamentals and


Applications, CRC Press, Boca Raton, 2008.

102. F. H. Quina, M. J. Politi, I. M. Cuccovia, et al., J. Phys.


Chem., 1983, 84, 361.

R. Zana, ed., Giant Micelles: Properties and Applications, CRC


Press, Boca Raton, 2007.

103. J. B. S. Bonilha, R. M. Z. Georgetto, E. Abuin, et al., J.


Colloid Interface Sci., 1990, 135, 238.

R. Zana, ed., Dynamics of Surfactant Self-Assemblies: Micelles,


Microemulsions, Vesicles and Lyotropic Phases, CRC Press,
Boca Raton, 2005.

104. D. Reichenberg, Ion-exchange selectivity, in Ion Exchange:


A Series of Advances, Vol. 1, ed. J. A. Marinsky, Marcel
Dekker, New York, 1966, p. 227.

M. Abe, Mixed Surfactant Systems, 2nd ed., CRC Press, Boca


Raton, 2004.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

Surfactant self-assembly
R. Zana, Gemini Surfactants: Synthesis, Interfacial and SolutionPhase Behavior, and Applications, CRC Press, Boca Raton,
2003.
D. Balzer and H. Luders, Nonionic Surfactants: Alkyl Polyglucosides, CRC Press, Boca Raton, 2000.
K. L. Mittal and D. O. Shah, eds., Adsorption and Aggregation of
Surfactants in Solution, CRC Press, Boca Raton, 2003.

23

J. Texter, ed., Reactions and Synthesis in Surfactant Systems, CRC


Press, Boca Raton, 2001.
A. Blume, J. Li, and T. Zemb, Self-assembly. Curr. Opin. Colloid
Interface Sci., 2007, 12, 99.
A. Blume, J. Li, and T. Zemb, Self-assembly. Curr. Opin. Colloid
Interface Sci., 2009, 14, 61.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc013

You might also like