You are on page 1of 33

Colloids and Surfactants

Assignment Title:
Micellization

Submitted To:

Dr Muhammad Usman
Submitted By:
M.Umair
Roll No:
526
Semester:
2nd Mphil (phi chem.)
Govt.College University Faisalabad

Q.No.1:
What is Micellization?
Ans:
One of the most characteristic properties of amphiphilic molecules is their capacity
to aggregate in solutions.The aggregation process depends, of course, on the
amphiphilic species and the condition of the system in which they are dissolved.
The abrupt change in many physicochemical properties seen in aqueous solutions
of amphiphilic molecules or surfactants with long hydrophobic chains when a
specific concentration is exceeded is attributed to the formation of oriented
colloidal aggregates. The narrow concentration range over which these changes
occur has been called the critical micelle concentration (CMC) and the molecular
aggregates that form above the CMC area are known as micelles. The difference
between micellar colloids and other colloids is that micellar colloids are in
dynamic equilibrium with monomers in the solution.

A micelle is an aggregate of surfactant  molecules dispersed in a liquid colloid. A


typical micelle in aquous solution forms an aggregate with the hydrophillic "head"
regions in contact with surrounding solvent, sequestering the hydrophobic single-
tail regions in the micelle centre.

This phase is caused by the physical behaviour of single-tail lipids in a bilayer. The
difficulty filling all the volume of the interior of a bilayer, while accommodating
the area per head group forced on the molecule by the hydration of the lipid head
group, leads to the formation of the micelle. This type of micelle is known as a
normal-phase micelle (oil-in-water micelle). Inverse micelles have the head groups
at the centre with the tails extending out (water-in-oil micelle).
Micelles are chemical structures formed with both hydrophilic (they’ll mix into
water) and hydrophobic (they don’t mix into water).

In the general case, micelles are formed when there is an ideal temperature in the
medium (called the Kraft temperature) and a certain concentration of electrolytes
(called the CMC(Critical Micelle Concentration) in the medium.

Micelles are how soaps work. Dirt usually is a form of an oil or some non-polar
compound, (Like a stain on your shirt). So, the hydrophobic end, (also non-polar),
adheres on to the dirt and the hydrophilic end, (being polar) bonds with the water.
(This is why you leave your dirty clothes in detergent-filled water for some time
and then scrub them, so that the above process can occur.)

Upon scrubbing the stain, you see the stain fading. This is because the micelles
(and the dirt) around the stain are pulled from the surface of the cloth and get
suspended into the water, which is then drained, leaving you with clean clothes.
The dirt gets attached to the end of the hydrophobic groups, i.e. the center.

Micelles are approximately spherical in shape. Other phases, including shapes such
as ellipsoids, cylinders, and bilayer, are also possible. The shape and size of a
micelle are a function of the molecular geometry of its surfactant molecules and
solution conditions such as surfactant concentration, temperature ,ph and ionoc
strength. The process of forming micelles is known as micellization.
Micelles are insoluble molecules (also known as lipid) when they goes in aqueous
solution like water they arrange themselves in spherical form (why spherical ?
because spherical is smallest geometry shape) .

It is a response to the amphipatic {word which means they have both hydrophobic
(water loving) and hydrophilic (water loving) } nature of fatty acids.

Micelles formed with fatty acids with usually one hydrophobic chain. the
hydrophobic parts come close to each other , there is a steric hindrance.
Q.No:2
Under what condition it takes place?
Micelles form only when the concentration of surfactant is greater than the critical
micelle concentration (CMC), and the temperature of the system is greater than the
critical micelle temperature, or Kraft temperature. The formation of micelles can
be understood using thermodynamics. Micelles can form spontaneously because of
a balance between entropy and enthalpy. In water, the hydrophobic effect is the
driving force for micelle formation, despite the fact that assembling surfactant
molecules is unfavorable in terms of both enthalpy and entropy of the system. At
very low concentrations of the surfactant, only monomers are present in solution.
As the concentration of the surfactant is increased, a point is reached at which the
unfavorable entropy contribution, from clustering the hydrophobic tails of the
molecules, is overcome by a gain in entropy due to release of the solvation shells
around the surfactant tails. At this point, the lipid tails of a part of the surfactants
must be segregated from the water. Hence, they start to form micelles. In broad
terms, above the CMC, the loss of entropy due to assembly of the surfactant
molecules is less than the gain in entropy by setting free the water molecules that
were "trapped" in the solvation shells of the surfactant monomers. Also important
are enthalpic considerations, such as the electrostatic interactions that occur
between the charged parts of surfactants.

The amphiphilic nature of surfactants gives rise to unusual solution properties


caused by the dual hydrophobic/hydrophilic character of the molecules. Ionic
surfactants behave as strong electrolytes in dilute solutions. On increasing the
concentration, the insubstantial balance between electrostatic and hydration
interactions get disrupted and the hydrophobic portions of the molecules attempt to
reorganize themselves in a way which will allow for a reduction of the unfavorable
hydrocarbon-water contact. When surfactants are dissolved in water, they initially
minimize their free energy by creating a monolayer on the air–water surface. In
this layer, the hydrophobic parts of the surfactants are directed towards less polar
air, while the hydrophilic groups are directed towards the polar water molecules.
As the surfactant concentration is increased, the surface becomes increasingly
populated by surfactant molecules, therefore, decreasing the surface tension of the
solution. On further addition, beyond a critical concentration (when the surface is
fully occupied), they start to aggregate leading them into a variety of structures in
which the hydrophilic head groups expose towards aqueous phase and hydrophobic
tails away from that.

The micelle formation is a dynamic process in which aggregates of approximately


similar size spontaneously begin to from over a narrow concentration range. The
size of the micelle is governed by effecting solution conditions also. The
aggregates formed are in dynamic equilibrium with dispersed monomer, a fact
which distinguishes micelles from other association colloids. Once micelles are
formed in solution they remain thermodynamically stable, with physic–chemical
properties distinct from those of the monomer solution.
Micelle formation and counter ions in the solution surrounding a charged surface
(if the surfactant is ionic) determine the rate at which the electrical potential will
change with the distance from the charged surface. For an understanding of the
electrical aspects of adsorption it is necessary to discuss the so–called electrical
double layer at interfaces. Hydrophobic cores of micelles have diameters of about
10–30 Å. The charged coat of ionic micelles, called the Stern layer, is 60–90٪
neutralized by counter ions in aqueous surfactant solution without added salt. The
core and the Stern layer form the ‘kinetic micelle’. Most of the remaining counter
ions are, however, located outside the shear surface in the region called ‘Gouty–
Chapman electrical double–layer’. According to Hartley model, the overall volume
of a micelle is approximately twice that of Stern layer. The nonionic micelles arrest
water molecules at the palisade layer (which includes the head groups and the first
few ethylene groups) by hydrogen bonding of water with the polyethylene oxide
groups. Water may remain trapped in this region.

Schematic representation of the regions of spherical micelle.

The surface charge of ionic micelles results in an electrical potential of the order of
100mV at the micelle–water interface with the same sign as the surfactant head
group. If a salt is added to the solution, the surface potential is partly neutralized,
decreases Coulombic repulsion between head groups, and allows the formation of
larger micelles. A solution having a single, very narrow distribution of micellar
size is often called monodisperse. As concentration of surfactant or salt or both in
water are increased, globular micelles gradually turn into larger rod– like micelles.
Under certain experimental conditions, spherical and rod–like micelles coexist in
the same solution; such systems are known as polydisperse. At higher
concentration of surfactants or salt, rod-like micelles begin to predominate and at
still higher concentration, lamellar liquid crystal phase may be formed.

The Gibbs energy of micellization then can be considered to be comprised of three


major contributions:

1. A suitable hydrophobic involvement arising from the transfer of the


hydrophobic moieties from water into the core of the aggregate,

2. A surface term which will account for the two opposing tendencies for the head
groups to crowd together in order to minimize water contact with the core of the
micelle, and to spread apart due to electrostatic repulsion, hydration, and steric
considerations, and

3. A packing term which requires water and hydrophilic head groups be expelled
from the interior of the aggregate, which will ultimately frontier the geometrical
handy forms available to aggregate.

Q.NO:3
Discuss Structure and Shape of Micelle and
what Factors do these depend?
Ans:
Micelles are lipid molecules that arrange themselves in a spherical form
in aqueous solutions. The formation of a micelle is a response to the amphipathic
nature of fatty acids, meaning that they contain both hydrophilic regions (polar
head groups) as well as hydrophobic regions (the long hydrophobic chain).
Micelles contain polar head groups that usually form the outside as the surface of
micelles. They face to the water because they are polar. The hydrophobic tails are
inside and away from the water since they are nonpolar. Fatty acids from micelles
usually have a single hydrocarbon chain as opposed to two hydrocarbon tails. This
allows them to conform into a spherical shape for lesser steric hindrance within a
fatty acid. Fatty acids from Glycolipids and phospholipids, on the other hand, have
two hydrophobic chains that are too bulky to fit into the a spherical shape as
micelles do.

Micelle in polar solvent(water) ( Some micelles have two hydrocarbon chain)


Micelle form spontaneously in water, as stated above this spontaneous arrangement
is due to the amphipatic nature of the molecule. The driving force for this
arrangement is the hydrophobic interactions the molecules experience. When the
hydrophobic tails are not sequestered from water this results in in the water
forming an organized cage around the hydrophobic tail and this entropy is
unfavorable. However, when the lipids form micelles the hydrophobic tails interact
with each other, and this interaction releases water from the hydrophobic tail and
this increases the disorder of the system, and this increase in entropy is favorable.

Structure and Shape:


The preferred structure of lipids in aqueous solutions are usually a bilayer sheet of
lipids rather than spherical micelles. This is because the two fatty acid chains are
too big and bulky to fit into the interior of a micelle. Therefore, micelles usually
have one hydrocarbon chain instead of two. Lipid bilayers" form rapidly and
spontaneously in an aqueous media and are stabilized by hydrophobic interactions,
Van der Waals attractive forces, and electrostatic interactions. The function of the
lipid bilayer is to form a barrier between the two sides of the membrane. Due to the
fact that the lipid bilayer consists of hydrophobic fatty acid chains, ions and most
polar molecules have trouble passing through the bilayer. The one exception to this
rule is water because water has a high concentration, small size, and a lack of a
complete charge. In order for a molecule to pass through the lipid bilayer it must
move from an aqueous environment to a hydrophobic environment and then back
into an aqueous environment.Therefore the permeability of small molecules is
related to the solubility of said molecule in a nonpolar solvent versus the solubility
of the molecule in water.
Micelles can also have a structure that is inside out of its normal structure. Instead
of having the hydrocarbon chains inside, they can face outside and while the polar
heads are arranged inside the sphere. This happens in a "water in oil" situation
because there is so much oil surrounding the drop of water that the hydrocarbon
chains face outside instead of inside.

In figure structure of micelle that will be formed when soap is dissolved in


hydrocarbon.

Micelle formation is caused by the molecules which have both hydrophilic and
hydrophobic part. Soaps are sodium or potassium salts of long chains of fatty acids
such that they have also a polar and a non-polar part. When dissolved in water the
polar part of the the soap molecule interacts with water molecules and the
hydrophobic part form a cluster as in the picture above. This type of structure is
called a micelle.

Soaps are generally made up of a hydrophobic part which is made up of the


hydrocarbon Chain and a metal which is normally (k) (Na) which I hydrophilic and
is attracted to water hence the structure is normally like this

RCOO– K +
Where R is the hydrocarbon Chain,K the Metal attached.

Types of micelles
When the concept of “Micelle” was introduced by McBain, it was rejected by the
Chairman of the Royal Society of London using two words: “Nonsense, McBain”.
It was a disheartening matter initially. However, the concept had subsequent
world–wide acceptance and its application in various fields of science and
technology witnessed its phenomenal importance. Over the last few decades,
studies of micellar systems occupy a distinct position in physical chemistry. The
scenario may be more appropriately described as a resurgence of the concept of
micelles.

(a) Normal micelle

Just above the CMC, the structure of a normal micelle can be considered as
roughly spherical. When the hydrocarbon portion of the amphiphile is a
hydrophobic chain, the micelle will consist of a liquid-like hydrocarbon core with
radius of roughly equal to the fully extended hydrocarbon chain length 12–30 Å).
The polar head groups with the surrounding water are roughly( arranged at the
micellar surface, which is rough.

The fluorescence and H-NMR measurements support the idea proposed by Menger
that water can penetrate inside the micelle up to a certain level. Partial molar
volume determinations indicate that the alkyl chains in the core are more expanded
than those in the normal liquid state.

(b) Reverse micelle

Amphiphiles in nonpolar solvents in presence of traces of water associate to form


the so–called reverse or inverted micelles. The structure of micelle is similar to that
of normal micelle but inverted, i.e., the polar head groups of the monomers are
present in the core and the hydrocarbon chains extending outwards into the solvent.

Reverse micelles represent one of the normal membranous structures in the cells.
The biological processes occurring in a reverse micellar system thus mimic those
involved in the in vivo membranous environment. For example,
AOT/isooctane/water reverse micelles have been used as a membrane–mimetic
system to study the requirement of tight bound lipids during the insertion of myelin
protein into the biomembrane.
Micelles with different sizes and properties can be made by changing the
water/surfactant ratio in the solution. Reverse micelles of this type have been
studied widely, primarily because of their usefulness as microreactors for chemical
and biochemical reactions.

(c) Mixed micelle

Mixtures of amphiphiles (specially, surfactants) have received wide attention for


several decades, both in theoretical studies and in practical applications, because of
their distinctive behavior in comparison with single amphiphile systems. In most
practical applications, mixtures of surfactants, rather than individual surfactants,
are used or purposely mixed to improve the properties of the final product. Mixing
of two or more surfactants in an aqueous solution leads to the formation of mixed
micelles.

The CMC of the mixed micelles fall within the highest and lowest individual
CMC values of components. In some cases, two surfactants interact in such a
fashion that the CMC of the mixture is always intermediate in value between those
of two pure components. In other cases, they interact in such a way that the CMC
of the mixture at some ratio of the two surfactants is less than either of the CMC.
When this situation arises, the system is said to exhibit synergism, the condition in
which the properties of the mixture are better than those attainable with the
individual components by themselves. For example, a long-chain amine oxide is
often added to a formulation based upon an anionic surfactant because the foaming
properties of the mixture are better than those of either surfactant by itself.

Binary mixtures of hydrocarbon surfactants have been studied to investigate the


micellar composition, aggregation number, molecular interactions, and free energy
of micellization, interfacial adsorption, micro– polarity employing a host of
techniques.
Factors affect the structure and shape of micelle:
The shape and size of the micelles can be controlled by changing chemical
structure of the surfactant as well as by changing the solution conditions such as
temperature, and electrolytes addition.

Effect of Temperatur:

The temperature effect varies the CMC value with the type of surfactant molecules.
The temperature has less effect on the micellar properties of ionic surfactants. This
is very well shown from the graph of sodium dodecyl sulfate(SDS) that CMC
versus temperature in Fig. The CMC varies in a irregular way by 10-20% over a
wide range. The shallow minimum around 25oC can be related with a similar
minimum in the dissolution of hydrocarbons in water. The decrease in solubility of
hydrocarbon increases the capacity of respective molecules of surfactant to
micellize.
(The CMC temperature dependence of Sodium Dodecyl Sulfate (top) and penta (ethylene
glycol) monodactyl ether.)

The polyoxyethylene non-ionic surfactants type deviate from this behavior and
show typically a monotonic, and much more pronounced decrease in CMC with
increase in temperature. Aqueous solutions of many non-ionic surfactants are
turbid at their fixed temperature that is known as the cloud point (The cloud point
of a fluid is the temperature at which solids dissolved are partially soluble, giving
precipitate of a second phase) . At temperatures above the cloud point, micellar
size increases and there is a corresponding decrease in CMC.

The Surfactants solubility can be strongly depend


on temperature
There are many important and intriguing temperature effects in surfactant self-
assembly. One, which is of great practical significance, is the dramatic
temperature-dependent solubility displayed notably by various ionic surfactant.
The solubility may be very low at cold temperatures and then increase by the order
of its magnitude in a relatively narrow temperature range. The phenomenon is
generally denoted as KRAFFT phenomenon with the temperature for the onset of
the strongly increasing solubility being known as the KRAFFT POINT or Krafft
temperature. The dependence of temperature on surfactant solubility in the region
of the Krafft point is illustrated in Fig. The Krafft point may vary dramatically with
slight difference in the surfactants chemical structure. Some general remarks can
be put for alkyl chain surfactant:

1. The Krafft point increases strongly as the length of alkyl chain increases. The
increase is not regular but displays an odd-even effect.

2. The Krafft point is strongly dependent on the counter ion and the head group.
But there are no general trends for the counter ion dependence. Salt addition
typically raises the Krafft point, while many other consolutes decrease it.

EFFECT OF CHEMICAL STRUCTURE

Some general remarks about the variation of the CMC with the chemical structure
of surfactant can be made:

1. The CMC decreases rapidly with increase in the alkyl chain length of the
surfactant. For eg. the ethylene oxide chain length of a non-ionic surfactant
increase which makes the molecule more hydrophilic and the CMC value
increases.
(The logarithm of the CMC amount linearly with the carbon atoms in the alkyl chain of the
surfactant. The slope is larger for a non-ionic surfactant or an ionic with added salt than for an
ionic surfactant without added electrolyte.)

(The logarithm of CMC (molar concentration) versus the carbon atoms in the alkyl chain for
octa(ethylene glycol) monoalkylethers ate different temperatures. From top to bottom, the
temperatures are 15.0, 20.0, 25.0, 30.0 and 40.0o C.)

2. Non-ionic surfactants mostly have low CMC values and have higher aggregation
numbers than their corresponding ionic counterparts of similar hydrocarbon chains.

3. A decrease in CMC, which for compounds with identical polar head groups is
represented by the linear equation: log [CMC] = A – Bnc where nc is the number
of carbon atoms in the chain and A and Bare constants for a homologous series.
4. Besides the major difference between ionics and non-ionics, the head group
effects are moderate. Cationics typically have slightly higher CMCs than anionics.
For non-ionics of the oxyethylene variety, a moderate increase in the CMC as the
polar head becomes larger.

5. The valency of the counter-ion is significant. While simple monovalent


inorganic counter ions give roughly the same CMC, increasing the valency to 2
gives a reduction of the CMC by roughly a factor of 4.

6. Micellar size for extremely cationic surfactant increases as thecounter-ion is


changing according to the series Cl− < Br− < I− , and for a distinctly anionic
surfactant according to the series Na+ < K+ < Cs+ .

7. Ionic surfactants includes organic counter-ions (e.g. maleates) have lower CMCs
and high aggregation numbers as compared to the inorganic counter-ions.

8. While alkyl chain branching and double bonds, aromatic groups or some other
polar character of the hydrophobic part produce sizeable turn in the CMC, a
considerable lowering of the CMC (one or two series of magnitude) results from
prefluorination of the alkyl chain. Partial fluorination interestingly may increase
the CMC, e.g. fluorination of the methyl group at terminal roughly doubles the
CMC value. The anomalous behavior of fractionally fluorinated surfactants is by
deleterious interactions of hydrocarbon and fluorocarbon groups.

ELECTROLYTE EFFECT

A most important matter is of added electrolyte effect on the CMC of ionics. This
is illustrated in Fig for the addition of 1:1 inert electrolyte to a solution of a
monovalent surfactant.
(Fig: Effect of sodium chloride addition on the CMC of different sodium alkyl sulfates. The dark lines are
predictions of electrostatic theory)

The addition of electrolyte to the solutions of ionic surfactants decreases the CMC
and the micellar size increases. This is because at high concentration of electrolyte
themicelles of ionic surfactants may become non-spherical. Salt addition gives a
considerable lowering of the CMC, which may amount to an order of magnitude.
The effect is moderate for short-chain surfactants but is much larger for long-chain
ones. As a consequence, at high concentrations of salt the variation of CMC with
the number of carbons in the alkyl chain is much stronger than without addition of
the salt. The rate of change at high salt concentrations becomes similar to that of
nonionics

Variation in Micelle Size and Structure

The driving force of micelle formation is the elimination of the contact between the
alkyl chains and water. The larger a spherical micelle, then the more efficient this
is, since the volume-to-area ratio increases. Decreasing the micelle size always
leads to an increased hydrocarbon-water contact. However, if the spherical micelle
was made so large that no surfactant molecule would reach from the micelle
surface to the centre, one would either have to create a void or some surfactant
molecules would lose the contact with the surface, introducing polar groups in the
center. Both alternatives are unsatisfactory.

We should note that the fact that the micelle radius equals the length of an
extended surfactant molecule does not mean that the surfactant molecules are all
extended. Only one molecule needs to be extended (in an all-trans state) to fulfil
the requirements mentioned, and the majority of the surfactant molecules are in a
disordered state with many gauche conformations. Spectroscopic studies have been
used to characterize the state of the alkyl chains in micelles in detail. This state
indeed is very close to that of the corresponding alkane in pure liquid oil.

At the surface of the micelle we have the associated counter ions, which in
number amount to 50-80% of the surfactant ions; as noted above, a number quite
invariant to the conditions. Simple inorganic counter ions are very loosely
associated with the micelle. The counter ions are very mobile and there is no
specific complex formed with a definite counter ion-head group distance. Rather,
the counter ions are associated by long-range electrostatic interactions to the
micelle as a whole. They remain hydrated to a great extent; cations especially tend
to keep their hydration shell.

Some water of hydration is thus accounted for by the associated counter ions and,
furthermore, the polar head groups are extensively hydrated. The micelle size, as
expressed by the radius of a spherical aggregated, may be obtained inter alia from
scattering experiments and from micelle self-diffusion. A related and equally
important characteristic of a micelle is the micelle aggregation number, i.e. the
number of surfactant molecules in one micelle. This is best determined in
fluorescence quenching experiments. To take an example, the aggregation number
of SDS micelles at is 60-70. The aggregation numbers deviating markedly from the
average the probability is very small.

Micellar morphology and micellar packing parameter:


The shapes of the micelles produced in aqueous media are of importance in
determining various properties of the surfactant solution, such as its viscosity, its
capacity to solubilize water–insoluble materials, and its cloud point. It is known
that the shape of micelles depends strongly upon the actual packing parameters in
the micellar assembly. The surfactant packing parameter, introduced by
Israelachvili, Mitchell, and Ninham, developed a general theoretical frame work
from which the secondary structures formed by a surfactant may be deduced based
on its molecular geometry.

The shape of the micelles is determined by the volume (v0) occupied by


hydrophobic group in the micellar core, the length of hydrophobic group in the
core (lc) and the cross sectional area (a0) occupied by hydrophilic group, which are
related to each other by packing parameter (p) p = v0/a0lc.
Relationship between the packing parameter and the shape of the
aggregate formed.

Effective shape of the Packing parameter Type of aggregation


surfactant molecule (p)
<1/3

cone spherical micelles


1/3–1/2

truncated cone wormlike micelles


1/2–1

bilayers
cylinder

Vesicles

>1

Reverse Micelle
Inverted cone
1.There are many factors that controlled the shape, size and CMC of micelles. The
temperature effect on the CMC varies with the type of surfactant molecules. CMC
for ionic surfactant varies in non monotonically, whereas for Non-ionic surfactants
it varies monotonic way that is decrease in CMC with increasing temperature.

2. The CMC decreases potently with increase in alkyl chain length of the
surfactant. Non-ionic surfactants have low CMC values and have higher
aggregation numbers than their ionic counterparts with similar hydrocarbon chains.

3. The addition of electrolyte to the solutions of ionic surfactants decreases the


CMC and increases the micelle size.

Q.No:4
What is Micellar Aggragation Number and what factors
does it depend?
Ans:
An aggregation number is a description of the number of molecules present
in a micelle once the critical micelle concentraton (CMC) has been reached. In
more detail, it has been defined as the average number of surfactant monomers in a
spherical micelle.
The aggregation number of micelles can be determined by isothermal
calarimetry when the aggregation number is not too high.
Another classical experiment to determine the mean aggregation number would
involve the use of a luminescent  probe, a quencher and a known concentration of
surfactant If the concentration of the quencher is varied, and the CMC of the
surfactant known, the mean aggregation number can be calculated.

The micellar aggregation number, which is the number of monomers making up


the micelles, is an important characteristic that determines the size and the
geometry of micelles. It is vital in determining the stability and practical
applications of investigated systems. The aggregation number depends on different
factors such as the nature of the surfactant, temperature, type and concentration of
the added electrolytes, organic additives,etc. The shape and size of the micellar
aggregates can, in principle, be determined by various methods such as viscosity,
light scattering, diffusion sedimentation velocity, sedimentation equilibrium,
ultrasonic absorption, time resolved fluorescence, small–angle neutron scattering
etc.

In a micellar solution, all micelles may not have the same aggregation and
polydispersity exists. However, for the sake of simplicity such polydispersity is
generally ignored for calculation purposes and only monodispersed micelles with
single aggregation number are taken into account.

The most common shape of micellar aggregates in solution is spherical, and hence
these are the most extensively studied. The main driving force for the self–
assembly of surfactant monomers into micelles is to minimize the hydrocarbon–
water contacts in solution. For this reason, the lower limit of the number of
surfactant monomers that form a micelle is dictated by the minimum number that
must come together to effectively shield one another from contact with water. The
very fact that discrete aggregates, typically containing on the order of 100
monomers or less, are observed in solution implies that there must exist a force
which opposes aggregate growth, or otherwise phase separation would be the
eventual result. In ionic surfactants, electrostatic repulsion between the ionic head
groups at the micellar surface provides the major contribution to this opposing
force. In the case of non–ionic surfactants, steric effect as well as a preference for
the hydration of the head group opposes micelle formation. Micelle formation,
therefore, represents a cooperative process, whereby a number of surfactant
monomers come together through a compromise of opposing forces. It is important
to note that micelles are not "monodisperse" in nature, i.e., they do not have a
uniform size of a fixed number of monomers. Rather there exists a distribution of
aggregate sizes from which the average number of monomers contained in a
micelle is taken as the mean aggregation number,Nagg.

5.What is critical micelle concentration (CMC)?


6.How CMC can be graphically identified??
7.How does CMC depends on various factors
Ans:
Critical micelle concentration:
The term CMC was established by Davis and Bury defining it as a concentration
range below which the surfactant molecules in the solution remain as monomers
and above which practically all additional surfactants added to the solution form
micelles. CMC is an important property of the surfactants which reflects its
micellization ability. A good surfactant will have a lower CMC value. Below the
CMC, the physicochemical properties of ionic surfactants resemble to those of
strong electrolytes and, above the CMC, these properties change dramatically,
indicating that a highly cooperative association takes place.

The CMC value of a surfactant micelle can be obtained by plotting an appropriate


physicochemical property versus the surfactant concentration and observe the
break point in the plot. The aggregation of surfactants/amphiphilic compounds can
be demonstrated by measuring solution properties such as surface tension, dye
solubilization, 1H–NMR, light scattering, fluorimetry, osmotic pressure, electrical
conductivity, ultrasound velocity, against the surfactant concentration. The value
of CMC depends on the measured solution properties and hence a difference in
their values is often associated with different experimental techniques. This is the
reason why a narrow concentration range is preferred for the CMC values.
At any temperature and pressure different experimental techniques may give
slightly different values for the CMC of a surfactant. However, Mukerjee and
Mysels in their vast compilation of CMC values, have noted that the majority of
values for a single surfactant are in good agreement and the outlying values are
easily accounted for. The CMC values are important in virtually all of the process
industry surfactant applications, from mineral processing to formulation of
personal care products and foods to drug delivery systems, and to new surfactant
remediation technologies. In these processes, surfactant must usually be present at
a concentration higher than the CMC because the greatest effect of the surfactant,
whether in interfacial tension lowering, emulsification, suspension stabilization, as
a delivery vehicle, or in promoting foam stability, is achieved when a significant
concentration of micelles is present.
Factors affecting the value of critical micelle
concentration
When the micelle formation takes place, the properties of solutions of surface–
active agents change markedly. Following are the main factors that are found to
affect the CMC in aqueous solutions.

(i) Structure of amphiphiles


(ii) Presence of additives

(iii) Experimental conditions such as temperature, pH, pressure, solvent, etc.

(1) Structure of amphiphiles

(a) The hydrophobic group

The CMC decreases strongly with increasing alkyl chain length of the surfactant.
While modification to the hydrocarbon chain (such as introducing branching, or
double bonds, or polar functional groups along the chain) usually leads to increase
in CMC, a dramatic lowering of the CMC (one or two orders of magnitude) results
from fluorination of the alkyl chain.

A generally used rule for amphiphile is that the CMC is halved by the addition of
one ethylene group to a straight chain hydrophobic group attached to a single
terminal hydrophilic group. The presence of branched chains or double bond
hinders micelle formation and thus increases the CMC. When the number of
carbon atoms in a straight–chain hydrophobic group exceeds 16, however, the
CMC no longer decreases so rapidly with increase in the length of the chain and
when the chain exceeds 18 carbon atoms it may remain substantially unchanged
with further increase in the chain length. This may be due to the coiling of these
long chains in water. A phenyl group that is a part of a hydrophobic group with
terminal hydrophilic group is equivalent to about three and one–half ethylene
groups. The replacement of hydrocarbon–based hydrophobic group by a
fluorocarbon–based one with the same number of carbon atoms appears to cause a
decrease in CMC. This is explained in term of the enhanced hydrophobicity.
(b) The hydrophilic group

Cationics typically have slightly higher CMC’s than anionics. The CMC’s of
nonionics are much lower than for ionics. For some nonionics, there is a moderate
increase of the CMC as the polar head becomes larger.

Effect of nature of polar group of ionic surfactants on the micellar properties has
been reported by Anacker and coworkers and they concluded that an important
factor controlling the micellar size was mean distance of closest approach of a
counterion to the charge center of the surfactant. Thus, for example,
decylammonium bromide forms very much larger micelles than
decyltrimethylammonium bromide because the Brcounterions are able to
approach more closely the charged nitrogen atom of decylammonium thus
effectively shielding the repulsive electrical forces and allowing larger micelles to
form. The more charged groups in the surfactants, the higher the CMC due to
increased electrical work required to form micelles. Zwitterionics appear to have
about the same CMC as ionics with the same number of carbon atoms in the
hydrophobic group. As the hydrophilic group is moved from the terminal position
to a more central position the CMC increases. It is because the hydrophobic group
seems to act as if it had become branched at the position of the hydrophilic group.

(ii) Presence of additives

(a) Effect of electrolyte


In solutions of increasing ionic strength, the forces of electrostatic repulsion
between head groups of ionic micelles are considerably reduced, enabling
micelles to form more easily, that is, at lower concentration. In other words,
addition of electrolyte to ionic surfactants causes CMC decrease [90], the effect
being more pronounced for anionic and cationic than for zwitterionic
surfactants, and more pronounced for zwitterionics than for nonionics. The
effect of the concentration of electrolyte is given by equation.
log CMC = – a log Ci + b
where a and b are constants for a given ionic head at a particular temperature
and Ci is the total (monovalent) counterion concentration in mole per dm3 .
The depression of the CMC in these cases is due mainly to the decrease in the
thickness of the ionic atmosphere surrounding the ionic head groups in the
presence of the additional electrolyte and the consequent decreased electrical
repulsion between them in the micelle. For sodium laurate and sodium
naphthenate, the order of decreasing effectiveness of the anion in depressing
the CMC is PO4 3– > B4O7 2– > OH– > CO3 2– > HCO3 > SO4 > NO3 >
Cl–.
For nonionics and zwitterionics, equation does not hold. Instead, the effect is
given by equation.
logCMC = –KCS + constant (CS < 1)
where K is a constant for a particular surfactant, electrolyte and temperature
and CS is the concentration of electrolyte in mole per dm3 .
The change in the CMC of nonionics and zwitterionics on the addition of
electrolyte has been attributed mainly to the “salting–out” or “salting– in” of
the hydrophobic groups in the aqueous solvent by the electrolyte, rather than to
the effect of the latter on the hydrophilic groups of the surfactant. Salting in or
salting out by an ion depends upon whether the ion is a water structure breaker
or a water structure maker.

(b)Effect of organic additives


Organic compounds affect the CMC either by penetrating into the micellar
region, or by modifying solvent–micelle or solvent–monomer interactions.
Non–polar compounds, such as hydrocarbons, that are believed to penetrate
into the inner portion of the core, decrease the CMC only slightly. Addition of
longer chain alcohols promotes micelle formation and lowers the CMC. The
magnitude of CMC decrease depends on the alkyl chain length of the organic
additive and the hydrophilic group associated with the chain. Urea, formamide,
and guanidinium salts are believed to increase the CMC of surfactants in
aqueous solution because of their disruption of the water structure. These water
structure breakers may also increase the CMC by increasing the entropy effect
accompanying micellization.

(iii) Experimental conditions


(a)Temperature
The CMC of amphiphiles in aqueous medium affected by the temperature in
such a way that firstly decrease with the increase in temperature to some
minimum, then increases with the increase in temperature.
The decrease in the CMC of the ionic surfactants with temperature increase at
lower temperatures is possibly due to dehydration of the monomers, whilst
further temperature increase causes disruption of the structured water around
the hydrophobic groups which opposes micellization.
Micelles of many polyoxyethylene nonionic surfactants increase rapidly in size
with rising temperature, whereas the micelle size of ionic surfactants decreases
with the increase in temperature . In some systems the rapid increase in micelle
size is noted only above a characteristic transition temperature and is
accompanied by an increase in micelle asymmetry. From viscosity data at
elevated temperature, it has been suggested that an increase in the extension of
the polyoxyethylene chains occur as the temperature is increased, resulting in
an increase in the amount of water physically trapped by the micelles and the
most drastic effect of temperature on nonionic surfactants is the effect on
solubility.

(b)pH

In amphiphiles bearing ionizable groups such as –NH2, – (CH3)2N–O and –


COOH, the degree of dissociation of the planar group will be pH dependent . In
general, the CMC will be high at pH values where the group is charged (low
pH from –NH2, and –(CH3)2N–O and high pH from –COOH) and low when
uncharged. Some zwitterionic surfactants become cationic at low pH, a change
that can be accompanied by a rapid rise in the CMC , or a more modest rise
depending on the structure and hence hydrophilicity of the zwitterionic form.

(c)Solvent

The polarity of medium favours surfactant association. Non–polar medium


offers environment similar to the surfactant tail so that their tendency of self–
association is reduced. It has been found that aggregation is retarded in some
organic solvents . CMC of surfactants were found to be lower in D2O than in
H2O because hydrophobic bonds may be stronger in D2O than in H2O.
Q.No:8

How thermodynamic parameters of micellization can be


calculated?

Ans:

Thermodynamic parameters of micellization


The process of micellization is one of the most important characteristics of
surfactant solution from theoretical as well as practical purposes and hence it is
essential to understand its mechanism (the driving force for micelle formation).
This requires analysis of the dynamics of the process (i.e., the kinetic aspects) as
well as the equilibrium aspects whereby the laws of thermodynamics may be
applied to obtain the free energy, enthalpy and entropy of micellization.

The pseudo-phase separation model which treats micelles as a separate phase


formed at and above the CMC, and the mass-action model which considers
surfactant monomer in solution to be in equilibrium with micelles of a fixed size
above the CMC. An extension of the mass-action model is the multiple equilibrium
model which consider the formation of aggregates of various sizes, accounting for
the observed polydispersity in aggregation numbers.

Phase-separation model
According to this model, micelles and counter ions are treated as separate phase.
However, the micelles do not constitute a “phase” according to the true definition
of this concept since they are not homogeneous and uniform throughout. Similarly,
there are problems associated with the application of the phase rule when
considering micelles as a separate phase.

Application of the phase-separation model to nonionic


surfactants:
To evaluate the thermodynamic parameters for the process of micellization a
primary requisite is to define the standard state. The hypothetical standard state
for the surfactant in the aqueous phase is taken to be the solvated monomer at unit
mole fraction with the properties of the infinitely dilute solution. For the
surfactant in the micellar state, the micellar state itself is considered to be the
standard state.
If µs and µm are the chemical potentials per mole of the unassociated surfactant in
the aqueous phase and associated surfactant in the micellar phase, respectively,
then since these two phases are in equilibrium
µs = µm

For non–ionized surfactant,

µs = µ ° s + RT ln as

where µ○ s is the chemical potential of standard state.

Since the micellar state is in its standard state

µm = µ ° m

At low concentration of free monomers, the activity as is replaced by mole fraction


Xs If the ΔG° mic is the standard free energy for the transfer of one mole of
surfactant from the solution to micellar phase, then

ΔG° mic = µ ° m – µ ° s = µm – ( µs – RT ln Xs) = RT lnXs


Eq.(a)

Assuming that the concentration of free surfactant in the presence of micelle is


constant and equal to the CMC value, XCMC, then

Δ G mic = RT ln XCMC Eq.(1)

XCMC is the CMC expressed as a mole fraction, therefore,

XCMC = ns / (ns + nH2O)

Since the number of moles of free surfactant, ns , is small compared to number of


moles of water, nH2O, therefore, Eq can be written as
XCMC = ns /nH2O

Substituting the value of this Eq. into the Eq. (1) and applying logarithm we get

ΔG° mic = 2.303RT (log CMC – log w)

where w is the number of moles of water (55.56 mol dm-3 at 25○C)

Application of Gibbs–Helmholtz equation to Eq. (1) gives

∂/∂T(ΔG° mic /T)p = – R(∂lnXCMC/∂T)P = ΔH ° mic /T2

Hence the standard enthalpy of micellization per mole of monomer,

Δ H mic , is ΔH ° mic = – RT2 (∂ln XCMC/∂T)P = R (∂ lnXCMC/∂(1/T))P

Also, standard entropy of micellization per mole of monomer, Δ S mic , is given


by

ΔS ° mic = (ΔH ° mic –ΔG° mic)/T

(b)Application of phase-separation model to ionic


surfactants
In the calculation of ΔG° mic, it is necessary to consider not only the transfer of
surfactant molecules from the aqueous to micellar phase but also the transfer of (1–
α) moles of counter ions from its standard state to the micelle. Therefore, Eq. (a)
can be written as

ΔG° mic = RT lnXs + (1–α) RT lnXx

where Xs and Xx are the mole fractions of surfactant ions and counter ions,
respectively.

The analogous equations for an ionic surfactant in the absence of added electrolyte
are

ΔG° mic = (2 – α) RT lnXCMC

ΔG ° mic = (2 – α) 2.303RT (log CMC – log w)


It is assumed that micellar phase is composed of the charged aggregates together
with an equivalent number of counter ions, and above Eqs. are approximated to

ΔG° mic = 2RT lnXCMC

ΔG ° mic = 4.606RT (log CMC – log w)

The enthalpy of micellization for ionic surfactants is given by

ΔH° mic = –2RT2 (∂lnXCMC/∂T)P

One of main criticism of phase separation model is that it predicts that the activity
of the monomer above the CMC remains constant. Surface tension and emf
measurements indicate decrease in monomer activity above CMC for ionic
surfactants.

(iii)Mass-action model
In the mass-action model, it is assumed that associated and unassociated surfactant
ions are in association-dissociation equilibrium and micellization is considered as a
reversible process. The mass action model was originally applied to ionic
surfactants but latter it was applied to nonionic surfactants also. Above CMC, the
concentration of monomer and micelle are interdependent due to equilibrium

n(S+ or S-or S) ↔(Sn - or Sn + ) Sn

where S+ or S- or S are surfactant monomer, Sn + or Sn - or Sn are surfactant


micelles, n is the aggregation number and Kmic is the micellization constant
(equilibrium constant)

Kmic = [Sn]/[S]n

At CMC, the free energy of micellization is given by

ΔG° mic = –RT lnKmic Eq(2)

At CMC, the aggregation number is a fixed quantity and the micellar concentration
at this stage is very low with respect to monomer, hence

ΔG° mic = –RT lnCMC


Applying Gibbs–Helmholtz equation and assuming the aggregation number to be
large and independent of temperature, the standard enthalpy of micellization can be
written as

ΔH° mic = –RT2 (∂ln Xcmc/∂T)P

The equation has a limited application when aggregation number varies with
temperature and concentration.

The micellization of ionic surfactants (Sor S+ ) along with the counter ions (I+ or
I- )

n(S+ or S- ) + m(I+ or I- ) ↔M(n-m)- or M (n-m)+

where n and m are aggregation number and number of counter ions that associate
with the ionic micelles.

Kmic = [M]/[S]n [I]m

The free energy of micelle formation per mole of monomer unit (ΔG 0 mic) is
then given by

ΔG ° mic = ΔG ° mic/n = {-1/(n ln[M]+ln[S]+m/n ln[I])}RT

At CMC, n is large. [S]= [I] = CMC (for a very small fraction of surfactant ions
from micelles), and

ΔG0 mic = – (1+ g1)RT lnCMC

where g1= m/n, is the fraction of counter ions bound to the micelle.

For nonionic surfactants, g1 = 0 and the equation reduces to Eq. (2). But when
counterion binding is 100%, i.e., g1=1,

ΔG° mic = –2RT lnCMC

The mass action model is more realistic model than the phase separation model in
describing the variation of monomer concentration with total concentration above
CMC. However, it suffers a serious limitation in that it considers monodispersity
of micelle size inspite of polydispersity. The phase separation model assumes
constant surfactant activity and hence surface tension above CMC although neither
of them remains constant. If aggregation number n is infinite then mass action and
phase separation models are equivalent.

Both the mass action and phase separation models, despite their limitations, are
useful representations of the micellar process and may be used to derive equations
relating the CMC to the various factors that determine it. Neither mass action nor
phase separation models are enough to explain the thermodynamics of
micellization completely. From the practical point of view a comprehensive
approach was developed, known as multiple equilibrium model which corrects the
flaws of mass action model.

Other thermodynamic models:


Another approach for that of small systems was formulated by Corkill and
coworkers and applied to systems of non–ionic surfactants. This multiple
equilibrium model considers equillibria between all micellar species present in
solution rather than a single micellar species, as was considered by mass–action
theory.

An interesting model of micelle formation based on geometrical considerations of


micelle shape has been proposed by Tanford. Equations are presented which relate
the micelle size and CMC to a size–dependent free energy of micellization. The
calculations are based on the assumptions of an ellipsoidal shape. The hydrophobic
component of the free energy is estimated in terms of the area of contact between
the hydrophobic core and the solvent. The hydrophilic component of ΔG0 mic, i.e.,
the free energy of repulsion between the head groups, is assumed to be inversely
proportional to the surface area per head group. This approach has been further
developed by Ruckenstein and Nagarajan and used in the prediction of the
properties of sodium octanoate micellar solutions.

(Note:Further two Questions in pdf file)

You might also like