01.06.2013 Views

Shampoo and Conditioner Science, Robert Y ... - Allured Books

Shampoo and Conditioner Science, Robert Y ... - Allured Books

Shampoo and Conditioner Science, Robert Y ... - Allured Books

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Practical Modern Hair <strong>Science</strong><br />

Chapter 3<br />

www.<strong>Allured</strong>books.com<br />

<strong>Shampoo</strong> <strong>and</strong><br />

<strong>Conditioner</strong> <strong>Science</strong><br />

<strong>Robert</strong> Y. Lochhead<br />

University of Southern Mississippi<br />

<strong>Shampoo</strong>s <strong>and</strong> conditioners are the highest volume of products<br />

sold in personal care. In this chapter, we will consider the science<br />

that underpins the functioning of these product types. The principal<br />

function of shampoos is to cleanse the hair. However, since the<br />

introduction of two-in-one shampoos in the 1970s, it has not<br />

been sufficient for a shampoo to merely cleanse the hair. Modern<br />

shampoos should at least cleanse, condition, make the hair easier<br />

to style, <strong>and</strong> fragrance the hair with a pleasant, lingering smell.<br />

Modern conditioners should lower the friction between hair fibers to<br />

allow easier grooming <strong>and</strong> alignment of the hair fibers while leaving<br />

them glossy <strong>and</strong> avoiding lankness.<br />

The science of shampoos <strong>and</strong> conditioners is still evolving <strong>and</strong><br />

in addition to describing fundamentals, this chapter attempts<br />

to take the reader to the frontiers of research in shampoo <strong>and</strong><br />

conditioner science.<br />

Introduction<br />

Located within the hair follicle is a sebaceous gl<strong>and</strong> that<br />

continuously excretes an oily material, known as sebum, onto<br />

the hair <strong>and</strong> scalp. This substance consists of compounds such as<br />

fatty acids, hydrocarbons, <strong>and</strong> triglycerides, <strong>and</strong> serves as nature’s<br />

conditioning treatment—providing lubrication <strong>and</strong> surface<br />

75


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

hydrophobicity, while potentially replenishing components of<br />

the cell membrane complex. However, after a day or so, buildup<br />

of this substance begins to result in a greasy look <strong>and</strong> feel.<br />

Moreover, particulate dust <strong>and</strong> dirt adhere readily to this sebum<br />

layer. In modern cultures such sebum-soiled hair is deemed to<br />

be undesirable, <strong>and</strong> therefore, it should be removed on a regular<br />

basis by a facile process. This process is, of course, shampooing.<br />

Sebum cannot be removed by water because oil <strong>and</strong> water do not<br />

mix. Aqueous shampoos can remove oily soil from the hair surface<br />

because shampoos contain surface-active agents, commonly<br />

abbreviated as surfactants. The molecules of these surface-active<br />

agents self-assemble into micelles, which are the agents that<br />

solubilize oily soils.<br />

To underst<strong>and</strong> how surfactants work, it is necessary to consider<br />

the exact process that leads to oil <strong>and</strong> water being incompatible.<br />

There are two different possibilities for substances to be insoluble<br />

in water. In one case, substances have stronger intermolecular<br />

cohesion than water. This is why substances like s<strong>and</strong>, clay, <strong>and</strong><br />

glass are insoluble in water; the molecules of s<strong>and</strong> attract each<br />

other more strongly than the molecules of water <strong>and</strong> this attraction<br />

leads to the s<strong>and</strong> being insoluble. This reason for the insolubility is<br />

exactly opposite to the reasons for the insolubility of hydrophobic<br />

substances such as oils. The intermolecular forces between the oil<br />

molecules are weaker than the intermolecular bonds between water<br />

molecules <strong>and</strong> the oils are expelled from water. This expulsion arises<br />

largely from entropy <strong>and</strong> the effect has been coined hydrophobic<br />

interaction. 1,2 From the time of the Phoenicians, it has been known<br />

that oil spreads to calm troubled waters. This effect arises from the<br />

fact that the spread oil has a lower surface tension than the water. At<br />

this point it is appropriate to consider the effect known as surface<br />

tension. Molecules in the bulk of liquids are attracted on all sides<br />

by their neighboring molecules. However, molecules at the surface<br />

are subjected to imbalanced forces because they are attracted by the<br />

underlying liquid molecules, but there is essentially no interaction<br />

with the vapor/gas molecules on the other side of the liquid/vapor<br />

76


Chapter 3<br />

boundary. This imbalance leads to a two-dimensional force at the<br />

surface, namely surface tension. The surface tension is numerically<br />

equal to the surface free energy. 3 The magnitude of surface tension<br />

directly correlates with the strength of the intermolecular forces.<br />

Water has hydrogen bonds, dipole-dipole interaction, <strong>and</strong> dispersion<br />

forces between its molecules, <strong>and</strong> as a consequence the surface<br />

tension of water is rather high—72 mN/meter at room temperature.<br />

On the other h<strong>and</strong>, only dispersion forces are present between the<br />

molecules of alkanes. As a consequence, the surface tension of<br />

alkanes is relatively low—ranging 20–30 mN/meter.<br />

Surfactants comprise molecules that contain two parts: a<br />

hydrophobic segment that is expelled by water <strong>and</strong> a hydrophilic<br />

segment that interacts strongly with water. Such molecules are said<br />

to be amphipathic (amphi meaning “dual” <strong>and</strong> pathic from the<br />

same root as pathos which can be interpreted as “suffering”). Thus,<br />

a surfactant molecule “suffers” both oil <strong>and</strong> water. This dual nature<br />

confers interesting properties on surfactants in aqueous solution.<br />

At very low concentrations, the surfactant is expelled to the surface,<br />

a process called adsorption. This adsorption causes the surfactant<br />

concentration at the surface to be much higher than the surfactant<br />

concentration in the bulk of the solution. At extremely low<br />

concentrations, when the surfactant molecules on the surface are<br />

located too far apart to effectively interact with each other, Traube’s<br />

Rule applies. Traube’s Rule states that the ratio of the surface<br />

concentration to the bulk concentration increases threefold for each<br />

CH 2 group of an alkyl chain. 4 This ratio is called the surface excess<br />

concentration. 5 According to this rule, soap with a dodecyl chain<br />

should have a surface excess concentration that is more than a halfmillion<br />

times its concentration in the bulk solution. At extremely<br />

low concentrations, the surfactant molecules on the surface act as a<br />

two-dimensional gas. As the concentration increases, the surfactant<br />

molecules begin to interact, but they are still mobile within the<br />

plane; they behave as two-dimensional liquids. At even higher<br />

concentrations, as the surfactant saturates the surface, the chains<br />

orient out of the surface plane <strong>and</strong> the chain-chain interactions<br />

77


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

cause the surfactant to behave as a two-dimensional solid. Irving<br />

Langmuir was awarded the 1932 Nobel Prize in Chemistry for<br />

measuring this effect <strong>and</strong> explaining it on a molecular basis. 6<br />

When a surfactant adsorbs to saturate an aqueous surface, the<br />

surface is largely composed of the surfactant’s hydrophobic groups;<br />

this means that the surface essentially has low surface energy. As a<br />

consequence of the low surface energy, the surface area is easier to<br />

exp<strong>and</strong> to a film. This means that the system is easier to foam, since<br />

aqueous foams really consist of water films with entrapped gas. If<br />

the foam surface is structured by the adsorbed surfactant, then foam<br />

stability can be achieved. 7<br />

Surfactant Micelles<br />

Relatively large aggregates form within solution just beyond<br />

the concentration at which the surface becomes saturated with<br />

surfactant. 8 These aggregates are surfactant micelles in which<br />

the hydrophobes are segregated within the core of the aggregate<br />

<strong>and</strong> the hydrophilic groups are located on the surface where they<br />

interact strongly with water. 9 For a given system, micelles initially<br />

form at the precise concentration at which the driving force for<br />

surface adsorption becomes equal to the driving force for aggregate<br />

formation. This driving force is the chemical potential of the<br />

surfactant species. The lowest concentration at which micelles form<br />

is named the critical micelle concentration (CMC). The aggregates are<br />

large; for example, micelles of sodium dodecyl sulfate at the CMC<br />

contain about 100 molecules <strong>and</strong> the thickness of the head group<br />

layer is about 0.4 nm. 10<br />

Surfactant micelles have liquid centers. They effectively solubilize<br />

hydrophobic substances only when the temperature of the system is<br />

above the Krafft point. Krafft found this phenomenon in 1895, <strong>and</strong><br />

68 years later Shinoda explained that the Krafft point corresponds to<br />

the melting point of the hydrated solid surfactant. 11<br />

Micelles have different shapes. The simplest shape is the<br />

spherical micelle that was postulated by Hartley in 1936. The<br />

shape of a micelle can be explained on the basis of the principle<br />

78


Chapter 3<br />

of opposing forces (see Figure 1). Two or three amphipathic<br />

molecules alone cannot form a stable micelle because micellization<br />

is essentially a cooperative process that requires the participation<br />

of many amphipathic molecules bound together by hydrophobic<br />

interaction. However, if hydrophobic interaction accounted solely<br />

for the formation of micelles, then the association would continue<br />

until phase separation occurred, as in oil separating from water.<br />

Therefore, there must be a force that opposes the hydrophobic<br />

association <strong>and</strong> controls the size of the micelles. This force is the<br />

repulsion between the head groups that could arise from ion-ion<br />

repulsion <strong>and</strong>/or hydration of the head groups. 12 Theoretically,<br />

the repulsive surface terms are difficult to h<strong>and</strong>le from a<br />

thermodynamic perspective but the presence of micelles has been<br />

validated experimentally.<br />

Figure 1. The shape of a surfactant micelle is determined by<br />

the balance between the mutual repulsion between hydrophilic<br />

groups at the micelle surface <strong>and</strong> the cohesion due to hydrophobic<br />

interaction. This has been dubbed the principle of opposing forces.<br />

If micelle structure was determined solely by thermodynamics,<br />

spherical micelles would always be favored over other shapes.<br />

However, real micelles are not restricted to a spherical shape;<br />

spherical structures account for only a small minority of micelles.<br />

The shapes of surfactant molecules <strong>and</strong> the way they can be packed<br />

79


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

also plays an important role in determining micelle shape. Although<br />

thermodynamics <strong>and</strong> packing geometries are inextricably linked,<br />

by considering the limits of possible packing arrangements we can<br />

obtain insight into the shapes of micelles <strong>and</strong> the transformation<br />

from one shape to another as physical <strong>and</strong> chemical conditions<br />

are changed. In this context, the many shapes of micelles, arising<br />

from the principle of opposing forces, can be appreciated by<br />

considering Packing Factor Theory (Figure 2). 13 First, consider a<br />

spherical micelle. In this instance the micelle radius, R, the volume<br />

of the hydrophobic core, v, <strong>and</strong> the surface area of the amphipathic<br />

molecule at the hydrophobe/water interface, a, are related by:<br />

80<br />

Eq. 1<br />

The radius of a micelle, R, cannot exceed the fully extended<br />

length, l, of the hydrophobe chain of the surfactant molecule. This<br />

gives the critical condition for the formation of spherical micelles:<br />

Eq. 2<br />

Figure 2. The packing factor of a surfactant molecule is the volume of<br />

the tail group divided by the volume of the cylinder subtended by the<br />

head group to the length of the tail group.


Chapter 3<br />

The fraction, v/al, is known as the packing factor (Figure 3).<br />

When the packing factor has a value of 1/3, the surfactant molecule<br />

can be approximated by a conical shape <strong>and</strong> the molecules pack into<br />

a sphere (Figure 4).<br />

Figure 3. Surfactant molecules with a packing factor of 1/3 have a<br />

shape that can be approximated by a cone.<br />

Figure 4. These conical molecules pack naturally into a sphere.<br />

When the packing factor has a value of ½, the micelles become<br />

cylinders (Figure 5), <strong>and</strong> when the packing factor has a value of<br />

1, the surfactant molecules pack as planar bilayers in a so-called<br />

lamellar structure (Figure 6).<br />

For ionic surfactants, the area per head group can be decreased<br />

by adding soluble salt to the solution to lessen the ionic repulsion<br />

between the head groups. (Salt also enhances the hydrophobic<br />

interaction. 14 ) Increase in salt <strong>and</strong>/or surfactant concentration causes<br />

spherical micelles to transition to rods <strong>and</strong> then to long worm-like<br />

micelles. 15 The wormlike micelles behave like polymers in solution. 16<br />

81


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

These micelles also form branched as well as linear structures, <strong>and</strong><br />

above a certain concentration (the critical overlap concentration, C*)<br />

they entangle just like polymer molecules 17 <strong>and</strong> display viscoelastic<br />

rheology. 18-20 This behavior is depicted in Figure 7 as it was<br />

explained by C<strong>and</strong>au in 1993. 21 An increase in salt concentration<br />

causes spherical or elliptical micelles to transition into rods, then<br />

to worms then to branched worms. As the surfactant concentration<br />

increases, the micelles form entangled networks. Consumers desire<br />

82<br />

Figure 5. Surfactant molecules with a packing factor<br />

of ½ pack naturally into cylinders.<br />

Figure 6. Surfactant molecules with a packing factor of 1 pack<br />

naturally into bilayer planes.


Chapter 3<br />

thicker shampoos, in part because they are easier to apply, but<br />

also for aesthetic reasons; a thicker formula is generally perceived<br />

as being more-luxurious. The desired rheology is achieved from<br />

formulations that contain worm-like micelles.<br />

Figure 7. Ionic surfactant micelles change shape as a function of ionic<br />

strength <strong>and</strong> surfactant concentration.<br />

Wormlike micelles do, however, show “non-polymeric” behavior<br />

at certain shear rates when the shear stress becomes independent of<br />

the shear rate <strong>and</strong> the relaxation time becomes monodisperse. 22 This<br />

behavior has been explained on the basis that the entanglements<br />

can be broken <strong>and</strong> reformed as the rod-like micelles disassemble<br />

<strong>and</strong> then reassemble upon passing through each other. 23-24 Systems<br />

like these have been dubbed “phantom networks” by Cates to<br />

signify that one micelle flows through another just as we imagine a<br />

phantom would pass through a wall. The phantom network behavior<br />

may explain why shampoos can show viscoelasticity without the<br />

“stringiness” observed in entangled polymer solutions.<br />

At higher concentrations, the rod-like micelles mutually repel,<br />

<strong>and</strong> this favors alignment into a nematic phase. At still higher<br />

concentrations the aligned rods pack in a hexagonal array to form<br />

hexagonal phase liquid crystals (Figure 8). The hexagonal phase<br />

has the properties of a clear ringing gel that is birefringent in<br />

polarized light.<br />

83


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

As the surfactant concentration is increased further <strong>and</strong>/or<br />

dissolved salt concentration is increased, the surface of the micelles<br />

becomes less curved until the large planar aggregates of the lamellar<br />

phase are formed (Figure 9). Modern shampoos consist essentially<br />

of entangled worm-like micelles <strong>and</strong> conditioners are usually in the<br />

form of the lamellar phase.<br />

In summary, shampoo <strong>and</strong> conditioner formulation essentially<br />

involves the preparation of surfactant mixtures that possess the<br />

84<br />

Figure 8. Rod-like micelles can pack into hexagonal liquid crystal phase.<br />

Figure 9. Increase in surfactant concentration causes micelles to transition from<br />

spheres to rods to hexagonal phase to lamellar phase to inverse hexagonal<br />

phase to inverse micelles.


Chapter 3<br />

aforementioned structures, while also being esthetically pleasing.<br />

The hair care formulation scientist has an ever-increasing variety<br />

of surfactants available in the formulation toolbox, <strong>and</strong> so these<br />

structures can be obtained via a wide range of concoctions.<br />

Nonetheless, attaining such stable structures is not a trivial task,<br />

due to the presence <strong>and</strong> interactions of so many ingredients in<br />

the typical formulation. Therefore, with historical knowledge<br />

involving many established ingredients already being relatively<br />

well-understood, it is a brave formulation chemist that opts to cut a<br />

new pathway. Moreover, it is also probably prudent to arrive at these<br />

structures in the most cost-effective manner. For these reasons, it<br />

is imperative to underst<strong>and</strong> how the surfactant structure, together<br />

with interactions with other molecules alters the nature of the<br />

aggregate structures.<br />

Oily Soil Removal Mechanisms<br />

The principal function of a shampoo is to remove oily soil from<br />

the hair. There are several principal detergency mechanisms for<br />

removing oily soils: “roll-up,” 25 emulsification, penetration, <strong>and</strong><br />

solubilization.<br />

In the roll-up mechanism, the detergent solution causes a steady<br />

increase in the contact angle of the oil at the oil/fiber/aqueous<br />

interface (Figure 10).<br />

Figure 10. In this mechanism the oil contact angle at the oil/water/fiber interface<br />

steadily increases until it “rolls up” <strong>and</strong> floats off of the solid surface. This<br />

mechanism was first reported by N. K. Adams.<br />

The oil droplet is rolled up on the surface, <strong>and</strong> when the contact<br />

angle reaches 180 degrees, the interfacial force that is holding it to<br />

the surface is overcome by the wetting tension of the oil <strong>and</strong> aqueous<br />

solutions on the fiber surface. Roll-up is favored by fibers that are<br />

85


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

oleophobic <strong>and</strong> hydrophilic. 26 The removal of oily soil by detergent<br />

compositions is not necessarily predictable due to the wide variation<br />

of the surface properties of hair that arise from prior treatments<br />

<strong>and</strong> weathering. Moreover, the transport of the detergent solution<br />

to the fiber surface can occur by three different routes: (i) along the<br />

fiber surface, (ii) through a previously applied permeable surface<br />

treatment, or (iii) through the body of the fibers (Figure 11).<br />

Roll-up of oily drops on fibers occurs when the contact angle<br />

exceeds a critical value <strong>and</strong> this causes the oily drop to adopt an<br />

unstable axially asymmetric attachment on one side of the fiber. 27<br />

The rate of roll up depends also on the viscosity of the oily soil,<br />

<strong>and</strong> mechanical action is often necessary to dislodge viscous oily<br />

soils from the fiber surface. In some cases, the oil forms a viscous<br />

emulsion when contacted by the detergent composition, <strong>and</strong> the<br />

resulting viscous soil can be difficult to remove from the fiber.<br />

“Perfect” hair is covered by a covalently attached monolayer of<br />

18-methyleicanosoic acid (18-MEA), which confers hydrophobicity<br />

on the hair. Modern grooming techniques <strong>and</strong> weathering removes<br />

this layer of 18-MEA. 28 Removal of the layer of 18-MEA results<br />

in hair becoming macroscopically hydrophilic. 29 The roll-up<br />

mechanism, therefore, should be expected to become more<br />

prominent on damaged rather than pristine hair.<br />

Initially if the fiber is completely coated in oil, or if the fiber itself<br />

is hydrophobic, the detersive solution cannot easily reach the oil/fiber<br />

interface, <strong>and</strong> the soil will be removed by emulsification (Figure 12).<br />

86<br />

Figure 11. In the roll-up mechanism, the detergent solution can be transported<br />

to the fiber/oil interface along the fiber surface, through a permeable coating<br />

on the fiber, or through the fiber itself.


Emulsification is favored by low oil/water interfacial tension that<br />

allows the oil surface to be exp<strong>and</strong>ed into an emulsion droplet. 30<br />

Figure 12. Emulsification can remove the soil if the interfacial tension between the<br />

oily soil <strong>and</strong> the surfactant solution is low.<br />

Chapter 3<br />

In the penetration mechanism of oily soil removal, surfactantrich<br />

phases penetrate the oil at the interface. This results in an<br />

interfacial liquid crystalline phase that swells <strong>and</strong> is broken off<br />

to reveal a fresh soil interface, <strong>and</strong> then the process is repeated<br />

again <strong>and</strong> again. 31 The penetration mechanism occurs with polar<br />

soils <strong>and</strong>/or phase separated coacervates of nonionic surfactants<br />

above the lower critical solution temperature (LCST). Spontaneous<br />

emulsification, in the absence of detersive surfactant, has been<br />

observed for non-polar-polar soil mixtures like sebum. 32 The<br />

penetration mechanism can occur with anionic surfactants that<br />

form coacervate phases in the presence of calcium salts. 33<br />

Solubilization is the process of incorporating a water-insoluble<br />

hydrophobic substance in the internal hydrophobic core of micelles.<br />

Direct solubilization can occur in the presence of an excess of<br />

surfactant micelles with respect to oily soil. 34 The rate of exchange<br />

of surfactant molecules between micelles is important because the<br />

micelles must re-assemble around the soil to solubilize the soil by<br />

encompassing it inside the micelle.<br />

Foam/Lather<br />

One essential attribute of a shampoo is its ability to produce<br />

a rich lather or foam. The important elements of a foam are<br />

the lamellae <strong>and</strong> the Plateau border. The micrograph in Figure<br />

13 depicts these structural features of a foam. The lamellae are<br />

stabilized by surfactants adsorbed at the air-water interface.<br />

87


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

Foams lose stability by two main mechanisms: draining of the<br />

liquid <strong>and</strong> puncture of the lamellae. The foam lamellae are the<br />

junctions between two foam bubble cells <strong>and</strong> the plateau border is<br />

situated at the triple-cell junction. The Laplace pressure in the liquid<br />

components of the foam is inversely proportional to the curvature<br />

of the interface. The higher curvature of the plateau border results<br />

in a lower pressure in that region <strong>and</strong> this causes the liquid in the<br />

foam to drain preferentially from the lamellae to the plateau borders.<br />

Based upon this reasoning, it can be understood that drainage can<br />

be hindered in two ways, namely by blockage of the lamellae or by<br />

blockage at the plateau border. About two decades ago, Des Goddard<br />

carefully measured the drainage from foam films <strong>and</strong> deduced<br />

that polyquaternium-24 adsorbed across the lamellar interface <strong>and</strong><br />

hindered the drainage of liquid from the foam. In addition, about<br />

thirty years ago, Stig Friberg concluded that certain liquid crystals<br />

blocked the plateau border region <strong>and</strong> delayed foam drainage <strong>and</strong><br />

conferred longer-term stability on surfactant foams. In the case of<br />

cationic polymers, hindered drainage of the lamellar liquid could be<br />

caused by adsorption of the cationic entities at the lamellar surface<br />

with the nonionic <strong>and</strong>/or anionic blocks in the lamellar liquid.<br />

88<br />

Figure 13. Micrograph showing surfactant foam structure.


Chapter 3<br />

Alternatively, formation of phase-separated coacervates between the<br />

cationic polymer <strong>and</strong> the anionic surfactant could result in blockage<br />

of the plateau border. Of course, if the interaction of the cationic<br />

polymer was strong enough to form “inverse micellar” structures,<br />

then there would be a possibility that the phase-separated particles<br />

could cause a local reversal of the curvature in the lamellae <strong>and</strong> this<br />

in turn would result in breakage of the lamellar film <strong>and</strong> subsequent<br />

foam destabilization. This type of foam destabilization mechanism<br />

has been extensively reported by Peter Garrett.<br />

Solid Foams<br />

Cationic conditioners<br />

that would normally be<br />

incompatible with liquid<br />

shampoos can be delivered<br />

from solid foams. Solid<br />

foams also make it possible<br />

to have one scent for the<br />

solid <strong>and</strong> then to allow<br />

a different fragrance to<br />

bloom when the solid is<br />

wetted by water. 35 The<br />

porous solids are made by<br />

mixing the surfactants,<br />

glycerin as a plasticizer,<br />

<strong>and</strong> water in the presence<br />

of a water-soluble polymer.<br />

Figure 14 shows a solid<br />

foam in which poly(vinyl<br />

alcohol) is the water-soluble<br />

polymer. After a heating Figure 14. Micrograph showing solid foam structure<br />

<strong>and</strong> mixing cycle, the (reproduced from US Patent Application 20110195098).<br />

porous solid is formed by<br />

aeration.<br />

89


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

The Anatomy of a <strong>Shampoo</strong> Formulation<br />

<strong>Shampoo</strong>s consist essentially of water, a primary surfactant,<br />

one or more co-surfactants, <strong>and</strong> soluble salt. Other ingredients<br />

are added for fragrance, preservation, conditioning, <strong>and</strong> styling<br />

attributes. Cleaning is achieved mainly by the primary surfactant,<br />

which is often an anionic surfactant that would adopt a conical<br />

shape if it was present in water alone. The co-surfactant is usually<br />

a nonionic or zwitterionic surfactant with a relatively small head<br />

group surface area. This molecular shape allows the co-surfactant<br />

to serve two roles: (i) it packs between the molecules of the primary<br />

surfactant to reduce the curvature <strong>and</strong> to promote the formation<br />

of worm-like micelles with their high viscosity <strong>and</strong> luxurious<br />

rheology; <strong>and</strong> (ii) it packs between the primary surfactant in the<br />

lamellae of the foam to provide good lather that is easily removed<br />

by rinsing. Salt enhances the function of the co-surfactant by<br />

“damping down” the ionic repulsion between primary surfactant<br />

head groups <strong>and</strong> promoting the formation of wormlike micelles. If<br />

excess salt or co-surfactant is added, shampoo compositions can<br />

separate into phases that contain co-existing micelles <strong>and</strong> liquid<br />

crystals. These phase-separated compositions often exhibit thin<br />

viscosities <strong>and</strong> haziness.<br />

The Primary Surfactant<br />

The lauryl sulfates have been the primary surfactant workhorses<br />

of the shampoo industry for decades. The sulfate head groups bear<br />

an anionic charge when dissolved in water. The long chain alkyl<br />

tail group has an average length of 12 carbon atoms. It is important<br />

to underst<strong>and</strong> that this is an average chain length; commercial<br />

lauryl sulfates have a distribution of chain length from as short as<br />

8 carbons to as long as 18 carbons. This chain length distribution<br />

changes from supplier to supplier <strong>and</strong> it also changes depending<br />

on the source of raw materials. Formulators should be aware that<br />

changes in the chain length distribution of the surfactants can lead<br />

to subtle changes in the properties of the shampoo.<br />

90


Chapter 3<br />

During the 1970s, triethanolamine lauryl sulfate was preferred<br />

as a primary surfactant due to its excellent cleaning properties <strong>and</strong><br />

luxurious flash foaming capability. However, it was replaced by<br />

laureth sulfates for two reasons: the concern over the formation of<br />

nitrosamines from secondary amine components <strong>and</strong> the reduced<br />

eye irritation exhibited by the laureth sulfates.<br />

Over the last two decades, the primary surfactants of most<br />

shampoos have been sodium laureth sulfate, ammonium lauryl<br />

sulfate, <strong>and</strong> sodium lauryl sulfate.<br />

The co-surfactant–often called the foam booster–has most<br />

prominently been selected from two types of materials: alkylamide<br />

MEA <strong>and</strong> alkylamidobetaines. Modern shampoos contain primarily<br />

betaines as co-surfactants.<br />

Enhancing Mildness<br />

Isethionates are surfactants noted for their mildness to skin,<br />

<strong>and</strong> for at least three decades, they have been the basis of non-soap<br />

detergent bars such as Dove (Unilever). They have been making<br />

inroads into shampoos based upon mildness claims. Moreover,<br />

Unilever researchers discovered that the mildness can be enhanced<br />

even further by including mildness benefit agents that can be<br />

flocculated by cationic polymers present in the formulation <strong>and</strong><br />

delivered as flocs upon dilution of the formulation. 36 The preferred<br />

benefit agent in this case is petrolatum; the cationic polymers<br />

are well known polymers like polyquaternium-10 <strong>and</strong> guar<br />

hydroxypropyltrimonium chloride. This could form the basis of<br />

shampoos that are mild to the skin.<br />

Certain non-cross-linked linear acrylic copolymers can lower<br />

the irritation potential of surfactants <strong>and</strong> provide products that are<br />

clear <strong>and</strong> highly foaming. 37 The preferred polymers interact with the<br />

surfactant <strong>and</strong> effectively shifting the CMC to higher concentrations,<br />

while lowering the critical aggregation concentration—the latter<br />

being the concentration at which the surfactant selectively interacts<br />

with the polymer rather than adsorbing at the liquid surface<br />

(Figure 15). It is postulated that free surfactant molecules <strong>and</strong><br />

91


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

free surfactant micelles are responsible for irritation of skin <strong>and</strong><br />

eyes <strong>and</strong> that binding of the surfactant to the polymer effectively<br />

reduces the concentration of free micelles. A measure of mildness<br />

is the delta CMC, which is defined as the difference between the<br />

CMC of the surfactant alone <strong>and</strong> the higher CMC of the surfactant<br />

in the presence of the polymer. Larger values of delta CMC for a<br />

particular surfactant are apparently correlated with lowering of the<br />

irritation potential. The delta CMC provides a measure that is useful<br />

for selecting, comparing, <strong>and</strong> optimizing polymers that reduce the<br />

irritation potential of selected surfactant systems. Carbomer <strong>and</strong><br />

acrylates copolymer have been identified as polymers that exhibit a<br />

satisfactory delta CMC.<br />

Conditioning <strong>Shampoo</strong>s<br />

Today’s conditioning shampoos are expected to confer wet-hair<br />

attributes of hair softness <strong>and</strong> ease of wet-combing, <strong>and</strong> the dry hair<br />

attributes of good cleansing efficacy, long-lasting moisturized feel,<br />

<strong>and</strong> manageability with no greasy feel.<br />

The origin of conditioning shampoos can be traced to the<br />

balsam shampoos of the 1960s followed by the introduction of<br />

polyquaternium-10 by Des Goddard 38,39 in the 1970s <strong>and</strong> 1980s in<br />

which he introduced the concept of polymer-surfactant complex<br />

coacervates that phase-separate <strong>and</strong> deposit on the hair during<br />

92<br />

Figure 15. Plot of surface tension vs. surfactant concentration for<br />

surfactant alone <strong>and</strong> for surfactant in the presence of polymer. The<br />

difference in the CMC induced by the presence of the polymer is<br />

claimed to be related to the effect of the polymer in enhancing the<br />

mildness of a shampoo.


Chapter 3<br />

rinsing. The first two-in-one shampoos depended on a complex<br />

coacervate being formed between anionic surfactant <strong>and</strong> the<br />

cationic hydroxyethylcellulose, polyquaternium-10. This complex<br />

was solubilized in excess surfactant <strong>and</strong> it phase-separated as a<br />

coacervate liquid phase upon dilution during the rinsing cycle.<br />

Later guarhydroxypropyltrimonium chloride was introduced as<br />

an alternative cationic polymer that worked on the same principle<br />

as polyquaternium-10. These two polymer types continue to<br />

dominate the compositions of conditioning shampoos. 40 Guar is a<br />

galactomannan <strong>and</strong> it is interesting that, in recent years, recently<br />

a new cationic galactomannan hydrocolloid, cationic cassia,<br />

has been claimed to confer conditioning shampoo benefits. 41,42<br />

Polygalactomannans consist of a polymannan backbone with<br />

galactose side groups. In guar gum, there is a pendant galactose<br />

side group for every two mannan backbone units. These galactose<br />

groups sterically hinder the substitutable C-6 hydroxyl unit,<br />

limiting the extent of possible cationic substitution on guar gum.<br />

In cassia, however, there is less steric hindrance of the C-6 hydroxyl<br />

group <strong>and</strong>, consequently, higher degrees of cationic substitution<br />

are possible with cassia (60% for cassia relative to 30% for guar).<br />

Cationic cassia can be used as a conditioning polymer in shampoos<br />

<strong>and</strong> conditioners to impart cleansing, wet-detangling, drydetangling,<br />

<strong>and</strong> manageability.<br />

The mechanism of conditioning shampoos depends upon the<br />

formation of polymer/surfactant coacervates that phase-separate<br />

during rinsing (Figure 16). Polyions in aqueous solution are<br />

surrounded by an electrical double-layer of counterions, <strong>and</strong> the<br />

location of the counterions with respect to the polyion is determined<br />

by a balance between chemical potential <strong>and</strong> electrochemical<br />

potential, called the Donnan Equilibrium. Surfactant ions contain a<br />

large hydrophobic group that makes them intrinsically less soluble<br />

in water than inorganic ions such as chloride or bromide. When<br />

surfactant ions interact with an oppositely charged polyion, they<br />

bind strongly <strong>and</strong> displace the water-soluble inorganic ions from<br />

the polyion; that is, they ion-exchange. Once the surfactant ions<br />

93


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

bind, hydrophobic interaction between the hydrophobic surfactant<br />

tails causes the polymer-surfactant complex to phase separate at<br />

concentrations below the surfactant critical micelle concentration.<br />

Above the CMC, the surfactant concentration is sufficiently high<br />

to form micelles or hemi-micelles along the polyion chain, <strong>and</strong> the<br />

polyion/surfactant complex is solubilized. Conditioning shampoos<br />

are formulated within the range of surfactant concentrations that<br />

correspond to this solubilized regime. When these shampoos are<br />

diluted to a concentration that is in the vicinity of the CMC, then the<br />

complex coacervate phase-separates. The separated phase is deposited<br />

on the hair during rinsing, <strong>and</strong> it can co-deposit other additives such<br />

as silicone conditioning agents or anti-d<strong>and</strong>ruff agents. Maximum<br />

coacervate deposition occurs at precise ratios of cationic polymer to<br />

anionic surfactant, but the optimum ratio for coacervation might not<br />

coincide with the best ratios for cleaning <strong>and</strong> foaming.<br />

Cationic guar has been a known additive for 2-in-1 shampoos<br />

for more than three decades. However, it has now been shown that<br />

improved post-shampoo detangling times are achieved by including<br />

a small degree of hydrophobic substitution in the cationic guar<br />

derivatives. 43<br />

94<br />

Figure 16. A schematic phase diagram that explains the mechanism of coacervate<br />

formation in 2-in-1 shampoos.


Chapter 3<br />

Synthetic copolymers of acrylamide <strong>and</strong> a Triquat monomer are<br />

postulated to provide improved deposition on hair <strong>and</strong> improved<br />

conditioning performance with respect to wet combing. 44<br />

Silicones have become st<strong>and</strong>ard ingredients in many conditioning<br />

shampoos for the smooth, silky hair feel that they confer. Silicones<br />

were introduced to shampoos as 2-in-1 conditioning agents in<br />

the 1980s. The introduction of silicones needed to overcome two<br />

substantial deficiencies: (i) silicones are well known defoamers, <strong>and</strong><br />

(ii) silicones are incompatible with typical shampoo compositions<br />

<strong>and</strong> they tend to separate due to their low specific gravity. Initial<br />

attempts to stably suspend the silicone included the use of water–<br />

miscible saccharides such as corn syrup. 45 Later products comprised<br />

xanthan gum in the shampoo as a suspending agent <strong>and</strong> acceptable<br />

foaming attributes were conferred on the shampoos by formulating<br />

with relatively high levels of alkyl sulfates as the primary surfactant,<br />

cocamide MEA as the co-surfactant, <strong>and</strong> ethylene glycol distearate<br />

as a surfactant structuring agent. 46 In the actual application, there<br />

is a technical contradiction involved in the deposition of silicone<br />

conditioning components from a detersive, cleansing system;<br />

the detersive system is designed to remove oil, grease, dirt, <strong>and</strong><br />

particulate material from the hair, <strong>and</strong> the conditioning agent<br />

has to be deposited on that same hair in one process. As a result,<br />

large excess amounts of silicone are used to ensure deposition,<br />

<strong>and</strong> one consequence of this is that large amounts of the expensive<br />

conditioning silicone can be rinsed away rather than deposited on<br />

the hair. Cationic polymer/anionic surfactant coacervates enhance<br />

the deposition of silicones on hair <strong>and</strong>, consequently, increase the<br />

efficiency of conditioning shampoos. 47,48<br />

Volatile cyclic siloxanes confer the desired silky initial feel, but<br />

these materials are difficult to formulate in consistent homogenous<br />

formulations, They tend to spread uncontrollably over the hair <strong>and</strong><br />

skin. 49 This effect can be controlled with polymeric silicone gels<br />

formed in volatile silicones to provide both the initial silky feel <strong>and</strong> a<br />

high viscosity <strong>and</strong> smooth feel when dry. 50 Branched molecules with<br />

a silicone core <strong>and</strong> hydrocarbon branches, or networks formed from<br />

95


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

these branched units, have been disclosed as suitable for improving<br />

sensory feel, while minimizing phase separation <strong>and</strong> conferring<br />

good shampoo removability. 51<br />

<strong>Shampoo</strong>s containing more than one cationic conditioning<br />

polymer <strong>and</strong> a quaternary silicone give more uniform deposition on<br />

hair than st<strong>and</strong>ard shampoos based on polyquaternium-10 as the<br />

sole conditioning polymer. Thus, a conditioning polymer “cocktail”<br />

comprising poly(acrylamide-co-acrylamidopropyltrimonium)<br />

chloride, guar hydroxypropyltrimonium chloride, <strong>and</strong> silicone<br />

quaternium-13 give uniform deposition on hair. In this instance, the<br />

claims are based upon multiple testing <strong>and</strong> analysis: 52<br />

96<br />

• A multiple attribute consumer assessment study that measured<br />

the attributes of cleanliness, wet-comb, dry-comb, hair softness,<br />

lather amount, <strong>and</strong> creaminess.<br />

• Secondary Ion Mass spectrometry to detect silicon on the hair<br />

surface. This method revealed that a st<strong>and</strong>ard commercial<br />

shampoo concentrated silicone on the cuticle edges of the hair,<br />

whereas the patent application shampoo “distributed silicone<br />

more evenly.”<br />

• X-ray photoelectron spectroscopy (XPS) to measure the thickness<br />

of the silicone polymer layer on hair from Si:C:O ratios. This<br />

method revealed that the commercial shampoo deposited<br />

a significant amount of silicone, <strong>and</strong> the patent application<br />

shampoo deposited only one or two molecular layers.<br />

• Instron ring compression as a measure of combability.<br />

Complex coacervates can also be formed from mixtures of<br />

cationic <strong>and</strong> anionic polymers. This could be the underlying<br />

mechanism in shampoos that include an anionic <strong>and</strong> cationic<br />

polymer that provide sleekness <strong>and</strong> gloss. 53<br />

Two drawbacks of silicones are that they often destabilize foam<br />

<strong>and</strong> the final compositions are hazy due to light scattered from<br />

the suspended silicone droplets. Initially, silicone copolyols were<br />

introduced to overcome the insolubility of silicones in shampoo<br />

compositions, but this drastically reduced the amount of silicone


Chapter 3<br />

deposited onto hair <strong>and</strong> compromised conditioning performance.<br />

Clear, silicone-containing conditioning shampoos have been<br />

formulated by adding trideceth-2 carboxamide MEA to reduce the<br />

silicone droplet size. 54 Transparent conditioning shampoos can be<br />

formulated by incorporating the silicone as microemulsified droplets,<br />

but the small microemulsion droplets tend to be rinsed away rather<br />

than deposited during the shampooing process. Moreover, coalescence<br />

of the droplets can lead to loss of transparency in the product during<br />

storage. Attempts have been made to overcome these challenges by<br />

including silicone emulsions with high internal viscosities, typically<br />

greater than 100,000 centistokes, but the high internal phase viscosity<br />

gives deposited silicone that is can be difficult to remove <strong>and</strong> this<br />

causes buildup with each consecutive shampooing. Such buildup<br />

usually reduces the volume of the desired hair style <strong>and</strong> causes<br />

“droop” <strong>and</strong> flatness. Fortunately, shampoo compositions providing<br />

superior conditioning to hair while also providing excellent storage<br />

stability <strong>and</strong> optionally high optical transparency or translucency can<br />

be obtained by combining low viscosity microemulsified silicone oil<br />

with cationic cellulose polymers <strong>and</strong> cationic guar polymers having<br />

molecular weights of at least about 800,000 <strong>and</strong> charge densities<br />

of at least about 0.1 meq/g. 55 Conditioning shampoo formulations<br />

that include a silicone microemulsion in a conditioning shampoo<br />

containing guar hydroxypropyltrimonium chloride <strong>and</strong> an anionic<br />

detersive surfactant have also been reported to be clear. 56,57 If pregelatinized<br />

starch, such as hydroxypropyl distarch phosphate, is<br />

included with polyquaternium-10, transparent conditioning shampoos<br />

can be obtained. 58<br />

Another way to minimize buildup is to treat the hair with waterin-water<br />

emulsions that can be prepared by including cationic<br />

polymers with soluble salts in surfactant compositions. 59 These waterin-water<br />

emulsions provide conditioning benefits with good spread<br />

of the conditioning phase on the hair <strong>and</strong> less chance of buildup.<br />

The living free radical polymerization techniques that have<br />

emerged in the last decade offer the prospect of preparing precise<br />

polymers with unprecedented accuracy in molecular structure <strong>and</strong><br />

97


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

variety of chemical types. 60 This technique has enabled synthesis of<br />

a wide diversity of block <strong>and</strong> graft polymers that were previously<br />

unattainable. Such polymers offer the prospect of conferring<br />

conflicting properties within one molecule, which in turn can lead<br />

to improved compatibilities in the same system while maintaining<br />

stability. These conflicting properties could possibly be achieved<br />

by blending different polymers, but different polymers do not mix<br />

readily at the molecular level <strong>and</strong> phase separation may result. 61<br />

Block, graft, <strong>and</strong> gradient copolymers serve to compatibilize such<br />

compositions <strong>and</strong> gradient polymers have been proposed for this<br />

purpose. Block copolymers comprising polycationic blocks <strong>and</strong><br />

nonionic blocks for surface deposition 62 <strong>and</strong> for improved foam<br />

retention 63 have been claimed, which are desired to deposit on<br />

hair in order to modify the chemical properties of the surface<br />

for protection or compatibility; to modify hair’s hydrophobic or<br />

hydrophilic surface properties; or to modify feel or mechanical<br />

properties of the substrate from two-in-one products. The polymers<br />

disclosed are block copolymers of polyTMAEAMS (methylsulfate<br />

[2-(acryloyloxy)ethyl]-trimethylammonium) g/mole) <strong>and</strong><br />

polyacrylamide.<br />

Conditioning shampoos can also be formulated to function<br />

by mechanisms other than cationic polymer-induced complex<br />

coacervation, such as:<br />

98<br />

• Conditioning can be achieved by including chain extended<br />

silicones in an anionic surfactant-based shampoo. Specific<br />

examples of useful silicones include silicone emulsions<br />

containing divinyldimethicone/dimethicone copolymer. 64<br />

• <strong>Shampoo</strong>s containing polyalkylene oxide alkyl ether particles<br />

give larger coacervate cohesive flocs (20–500 microns) that<br />

resist shear <strong>and</strong> confer superior deposition efficiency on hair for<br />

good wet conditioning. 65<br />

• Conditioning shampoos containing a polyester formed from<br />

adipic acid <strong>and</strong> pentaerythritol provides conditioning for dry<br />

hair (possibly from reduced hair friction), with no greasy feel. 66


• Inclusion of polybutene is thought to increase the deposition<br />

of silicone conditioners <strong>and</strong> provide improved conditioning<br />

benefits, such as wet <strong>and</strong> dry feel <strong>and</strong> combing.<br />

Chapter 3<br />

O’ Lenick disclosed a unique class of alkyl polyglucoside quaternary<br />

surfactants possessing all the multifunctional attributes of cleansing,<br />

conditioning, <strong>and</strong> self-preserving. 67 This could have the potential of<br />

greatly simplifying the formulation of multifunctional shampoos.<br />

A conditioning shampoo that contains a conditioning gel<br />

phase in the form of vesicles is described by Unilever researchers. 68<br />

Cationic conditioners are usually incompatible with anionic<br />

shampoos, <strong>and</strong> consequently conditioners based upon cationic<br />

surfactants are usually applied as separate post-shampoo products.<br />

The Unilever researchers prepared a conditioning gel phase by<br />

combining a small amount of water, fatty alcohol, a long-chain<br />

secondary anionic surfactant (sodium cetostearyl sulfate), <strong>and</strong><br />

a long-chain cationic surfactant (behenyltrimethylammonium<br />

chloride), <strong>and</strong> subjecting the mixture to high shear to form a stable<br />

vesicular gel phase. Prolonged shear causes the lamellar gel phase to<br />

roll-up into an array of multilamellar vesicles (Figure 17). The gel<br />

phase was added to a dilute primary surfactant solution (sodium<br />

laureth sulfate) to form a conditioning shampoo that conferred good<br />

wet smoothness on hair.<br />

Figure 17. Lamellar gel subjected to high shear rolls up into vesicles of gel phase that can<br />

be used for conditioning. (Figure reproduced from US Patent Application 20110243870).<br />

Deposition of Particles on Hair to Confer Styling Benefits<br />

Whereas conditioning shampoos are formulated to reduce hair<br />

inter-fiber friction, some consumers need an increase in friction in<br />

99


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

order to achieve styling benefits. Factors that influence hair body<br />

<strong>and</strong> fullness include hair diameter, hair fiber-to-fiber interactions,<br />

natural configuration (kinky, straight, wavy), bending stiffness,<br />

hair density, <strong>and</strong> hair length. Increases in friction can be achieved<br />

by depositing particles such as titanium dioxide, clay, pearlescent<br />

mica, or silica on the hair surface. Particles can be deposited for<br />

more purposes than merely increasing inter-fiber friction, e.g., for<br />

conferring color, for slip (spherical particles are best for this), <strong>and</strong><br />

for conditioning (hollow silica, hollow polymer particles). Hollow<br />

particles can be included in shampoo to increase hair volume. 69,70<br />

Deposited hollow particles that can increase fiber-fiber interaction<br />

include complexes of gas-encapsulated microspheres (such as silica<br />

modified ethylene/methacrylate copolymer microspheres <strong>and</strong><br />

talc-modified ethylene/methacrylate copolymer microspheres);<br />

polyesters; <strong>and</strong> inorganic hollow particles.<br />

It has already been noted that cationic guar enhances<br />

the deposition of conditioning agents. In a like manner, this<br />

macromolecule enhances the deposition of particles on hair. 71<br />

Silicones <strong>and</strong> particulates can be deposited simultaneously. Thus,<br />

enhanced deposition of particulate actives, such as zinc pyrithione<br />

(shown on cadaver skin treated in a Franz diffusion cell), has<br />

been reported 72 from shampoos comprising a water-soluble<br />

silicone (such as silicone quaternium-13, cetyltriethylammonium<br />

dimethicone copolyol phthalate, or stearalkonium dimethicone<br />

copolyol phthalate), a cationic conditioning agent (such as<br />

acrylamidopropyltrimonium chloride/acrylamide copolymer, or<br />

guar hydroxypropyltrimonium chloride), a cleansing detergent,<br />

<strong>and</strong> suspending agents (such as carbomer, hydroxyethylcellulose,<br />

<strong>and</strong> PVM/MA decadiene cross-polymer) to insure homogeneous<br />

distribution of the insoluble active.<br />

Hydrophobic modification of cationic hydroxyethylcellulose<br />

is claimed to endow better efficacy. Thus, polyquaternium-24,<br />

a hydrophobically modified cationic hydroxyethylcellulose, is<br />

also disclosed as being a preferable thickener for zinc-depositing<br />

compositions. 73<br />

100


Chapter 3<br />

It has been discovered that responsive particles, with two<br />

contrasting polymers adsorbed to the particle core, 74 can adsorb<br />

to the hydrophilic hair surface <strong>and</strong> render it hydrophobic, thereby<br />

conferring conditioning attributes to the hair. For example, grafting<br />

of aminopropyl-terminated dimethicone <strong>and</strong> polyethylenimine<br />

on titanium dioxide particles produces responsive particles.<br />

These particles form stable dispersions in water <strong>and</strong> aqueous<br />

solutions because they are sterically stabilized by expansion of<br />

the polyethylenimine into the aqueous medium. However, when<br />

they are deposited on hair <strong>and</strong> dried, the polyethylenimine layer<br />

collapses <strong>and</strong> the dimethicone layer exp<strong>and</strong>s to render the surface<br />

hydrophobic. The usefulness of these responsive particles is<br />

demonstrated by including them in typical conditioning shampoo<br />

<strong>and</strong> conditioner formulations. In the case of the shampoo, inclusion<br />

of the responsive particles results in a higher water contact angle on<br />

the treated hair <strong>and</strong> the conditioner with particles causes an increase<br />

in the hydrophobicity of the hair. On the other h<strong>and</strong>, shampoos<br />

containing ethoxylated alcohols have been found to enhance the<br />

deposition of large particle silicones (5–2000 microns), <strong>and</strong> in this<br />

case it is claimed that cationic polymer is not required. 75<br />

Two-phase Systems for Visual Attributes:<br />

There is esthetic appeal to products that exist as separate phases<br />

in the bottle but which mix during application to provide added<br />

benefit, such as moisturizing or conditioning, by interaction of the<br />

components of the two phases. The most obvious way to formulate<br />

such products is to use the immiscibility of water <strong>and</strong> oil in<br />

formulations that are shaken prior to use to produce a metastable<br />

emulsion. However, when a surfactant is included in the system such<br />

a visually attractive phase separation can be mixed into an emulsion<br />

due to shear in manufacturing <strong>and</strong> packing operations. There are<br />

known de-emulsifiers, which are widely used in the oil industry,<br />

but these demulsifiers also tend be defoamers that compromise the<br />

lather of shampoos. Neutralized polyacrylate can be added as a nonemulsifying<br />

foam stabilizer to yield phase-separated compositions<br />

101


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

that resist the production difficulties to make phase-separated<br />

shampoos that form temporary emulsions upon shaking <strong>and</strong> foam<br />

during use. 76 A two-phase shampoo system can also be formed by<br />

mixing polar lipophilic shampoo components with non-polar lotion<br />

constituents such as mineral oil. 77<br />

Under appropriate conditions, phase-separated systems can be<br />

prepared from polymer solutions or micellar surfactant solutions. If<br />

two distinct aqueous phases are desired in a composition, one must<br />

take into consideration the thermodynamics of coexisting phases<br />

<strong>and</strong> the driving force for such phase separation that comes directly<br />

from the chemical thermodynamics of the system. This is especially<br />

the case for systems that contain polymers or micelles because the<br />

configurational entropy is reduced as molecules are assembled<br />

into polymers or aggregated into micelles, <strong>and</strong> mixing can become<br />

unfavorable. If the free energy of mixing is insufficient to maintain<br />

uniform dispersion, spontaneous phase separation will occur.<br />

Phase separation becomes more likely as the micellar aggregates or<br />

polymers get bigger. The addition of salts to ionic surfactant micellar<br />

systems causes a reduction in the surfactant intra-micellar headgroup<br />

interaction, <strong>and</strong> often an increase in hydrophobic interaction.<br />

This can cause a pronounced increase in micelle size <strong>and</strong> consequent<br />

phase separation into a surfactant-rich phase <strong>and</strong> a surfactant-poor<br />

phase. This approach has been adopted by adding mineral salt to<br />

induce two distinct layers, 78 <strong>and</strong> by adding the detergent builder,<br />

sodium hexametaphosphate, to cause phase separation. In this case,<br />

a thickener is required <strong>and</strong> the system comprises a surfactant, a<br />

thickener, a polyalkylene glycol, <strong>and</strong> a non-chelating mineral salt.<br />

The system spontaneously separates into two layers.<br />

A multiphase composition comprising surfactant, betaines,<br />

co-surfactant (such as an alkyl ether carboxylate, an acylglutamate;<br />

or an acylisethionate), <strong>and</strong> an appropriate concentration of salt<br />

forms a stable multiphase system that becomes temporarily<br />

uniformly dispensed upon agitation. 79<br />

Multiphase cleansing products have been introduced that go<br />

beyond mere phase separation insofar as the separate phases can be<br />

102


Chapter 3<br />

arranged to form visually attractive patterns inside a transparent<br />

container. 80 The phases comprise an aqueous cleansing phase, a<br />

benefit phase, <strong>and</strong> a non-lathering structured phase. The aqueous<br />

cleansing phase must be capable of adequate lathering. 81 The<br />

benefit phase comprises hydrophobic component(s) or conditioning<br />

components. These products are designed at the nanoscale: the<br />

structured phase can be a lamellar-phase formed by adding sufficient<br />

electrolytes to an appropriate surfactant. Structurants such as<br />

starch have been used in personal cleansing formulations, 82 but the<br />

surfactant itself can be structured. Thus, lamellar phase does exhibit<br />

a yield stress that is sufficient to stably suspend the benefit phase.<br />

However, the yield stress of lamellar phase can vary dramatically<br />

with temperature, <strong>and</strong>, in order to overcome this problem, the<br />

cleansing <strong>and</strong> benefit phases were density matched by adding<br />

microsphere particles to reduce the specific gravity of the cleansing<br />

phase or high density particles to the benefit phase to increase its<br />

specific gravity. In this context, it is interesting that it has been<br />

recently disclosed that controlled phase separation <strong>and</strong> deposition<br />

could conceivably be achieved by loading the desired “active” phase<br />

into hollow-sphere polymer carriers, 83 <strong>and</strong> again it has been reported<br />

that certain cationic guar derivatives can enhance the deposition of<br />

conditioning additives <strong>and</strong>/or solid particle benefit agents. 84<br />

Lamellar phase, especially if it is made from unneutralized longchain<br />

fatty acids, usually displays poor dispersion kinetics <strong>and</strong> a<br />

lather that is slow to build up or slow to rinse off. However, it has<br />

surprisingly been discovered that swollen lamellar gels can exhibit<br />

both high product viscosity <strong>and</strong> fast dispersion kinetics if they are<br />

formed by combining C16-24 normal monoalkylsulfosuccinates<br />

with n-alkyl fatty acids of approximately the same chain length. 85 In<br />

this context, Guth claimed a composition that was low-irritating to<br />

skin <strong>and</strong> eyes but synergistic in foaming by combining zwitterionic<br />

surfactants-fatty acid complexes with sulfosuccinates, 86 <strong>and</strong> Pratley<br />

reported synergistic foaming <strong>and</strong> mildness from compositions<br />

with combinations of specific long-chain surfactants with specific<br />

short-chain surfactants <strong>and</strong> these included fatty acids <strong>and</strong><br />

103


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

alkylsulfosuccinnates. 87 Amine-oxide copolymers have also been<br />

claimed as suds-enhancers. 88<br />

Concentrated Cleansing Compositions<br />

Most personal care products are based on aqueous compositions<br />

but concentrated cleansing/personal care compositions offer<br />

the benefits of lower transportation costs, less packaging, <strong>and</strong><br />

convenience for air travelers. If these products are solid, they<br />

must possess sufficient strength to resist the forces of extrusion<br />

during manufacture, shipping, <strong>and</strong> h<strong>and</strong>ling, but should disperse<br />

rapidly in water during use. Porous solid particles that are<br />

strengthened by hydrophilic polymers such as poly(vinyl alcohol)<br />

or hydroxylpropylmethylcellulose have been shown to exhibit the<br />

desired properties. 89 Control of interconnectivity of the porous<br />

network is vital to this application <strong>and</strong> is described by a star volume,<br />

a structure model index, or a percent open cell content.<br />

<strong>Conditioner</strong>s<br />

Conditioning of damaged hair is commonly achieved by<br />

treatment with aqueous formulations that contain fatty alcohols,<br />

cationic surfactants, <strong>and</strong> (optionally) silicones. These components<br />

are considered to adsorb in a hydrophilic head-down, hydrophobic<br />

tail-up conformation that confers hydrophobicity on the damaged<br />

hydrophilic hair surface. The role of a conditioner is to confer sleek<br />

lubricity <strong>and</strong> gloss on the hair. <strong>Conditioner</strong>s are usually based<br />

upon cationic surfactants, <strong>and</strong> they most often are in the form<br />

of emulsions of multi-lamellar vesicles. <strong>Conditioner</strong>s comprise a<br />

primary cationic surfactant, a co-surfactant, <strong>and</strong> dissolved salt.<br />

Conventional conditioner formulations are based upon lamellar<br />

gels or emulsions using either ceto-stearyl trimethylammonium<br />

chloride or distearyldimethylammonium chloride as cationic<br />

surfactants <strong>and</strong> ceto-stearyl alcohol as co-surfactant.<br />

As a primary surfactant, the vast majority of conventional<br />

conditioners contain either cetyl/stearyl trimethylammonium<br />

104


Chapter 3<br />

chloride or distearyldimethylammonium chloride. The secondary<br />

surfactant is most often ceto-stearyl alcohol.<br />

Cetyl/stearyl trimethylammonium chloride is a conical molecule<br />

according to Ninham’s packing factor. On the other h<strong>and</strong>, cetostearyl<br />

alcohol consists of molecules with the approximate shape<br />

of an inverted conical molecule. The role of ceto-stearyl alcohol<br />

in a conditioner is to pack between the cationic cones <strong>and</strong> convert<br />

the micellar structure into a lamellar structure with just enough<br />

curvature to form a vesicle. Distearyldimethylammonium chloride<br />

spontaneously forms vesicles in the presence of salt, <strong>and</strong> therefore<br />

there is usually no need to add a long-chain alcohol to conditioner<br />

formulations based upon distearyldimonium chloride.<br />

These products form a gel matrix that confers conditioning<br />

benefits from rinse-off products. They have been the basis of hair<br />

conditioners for the last half-century, <strong>and</strong> they provide excellent<br />

detangling, wet- <strong>and</strong> dry-combing, <strong>and</strong> good anti-static properties,<br />

but they can leave the hair feeing lank <strong>and</strong> greasy, <strong>and</strong> they give a<br />

long-lasting slippery feel during rinsing which is perceived by some<br />

consumers as an unclean hair feel.<br />

Pristine hair, as it emerges from the scalp, is coated with a<br />

covalently bound layer of 18-methyleicanosoic acid (18-MEA). 90-92<br />

It has been shown that the layer of 18–MEA confers hydrophobicity<br />

<strong>and</strong> boundary lubrication on hair fibers. 93 This discovery has<br />

influenced researchers to seek to include 18-MEA in conditioner<br />

formulations. 94 Pristine hair shows a measured advancing water<br />

contact angle that is high, but a receding contact angle that is<br />

likewise high, <strong>and</strong> the hair tends to align. However, once the<br />

18-MEA layer is removed, the receding contact angle is low (even<br />

approaching 0 degrees), <strong>and</strong> this corresponds to cuticle edges that<br />

are essentially hydrophilic. This means that the major differences<br />

for such 18-MEA deficient hair would be in its drying behavior<br />

rather than its wetting characteristics. The low receding contact<br />

angle would tend to “pin” the water to the hair. This would<br />

lead to longer drying times during which the capillary forces<br />

imparted by the water between hair fibers would tend to cause<br />

105


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

the hair fibers to clump <strong>and</strong> entangle. The inclusion of 18-MEA<br />

in prototype conditioner formulations that were based upon<br />

stearoxypropylmethylamine, dimethylaminopropylstearamide, <strong>and</strong><br />

stearyltrimethylammonium chloride left the conditioner on the<br />

hair surface, <strong>and</strong> this in turn yielded improvements in inter-fiber<br />

lubricity due to improved deposition at the hair surface.<br />

Conditioning Polymers in Hair Straightening Applications<br />

The two main processes for relaxing or straightening hair are<br />

hair treatment with a reducing agent to cleave the disulphide cystine<br />

bridges (S—S) within the hair structure, <strong>and</strong> treatment of stretched<br />

hair with a strong alkaline agent.<br />

Repeated relaxation treatments can cause significant hair damage,<br />

to both the cuticles <strong>and</strong> the cortex. The damage can be assessed by<br />

measuring the porosity of the hair, <strong>and</strong> the porosity of the keratin<br />

fibers can be measured by fixing 2-nitro-para-phenylenediamine<br />

at 0.25% in an ethanol/buffer mixture (10:90 volume ratio) at pH<br />

10 at 37°C for 2 minutes. Cationic <strong>and</strong> amphoteric polymers, such<br />

as polyquaternium-6, polyquaternium-7, <strong>and</strong> polyquaternium-39,<br />

added to hair relaxer formulations, mitigate this degradation of the<br />

hair structure. Also, the inclusion of high molecular-weight (>106<br />

g/mole) copolymers of acrylamide <strong>and</strong> diallyldimethylammonium<br />

chloride, acryloyloxytrimethylammoniumchloride, or<br />

acryloyloxyethyldimethylbenzylammonium chloride in the relaxing<br />

formula results in significant reduction in the hair structural<br />

damage caused by alkaline relaxation.<br />

Conditioning Polymers<br />

Cationic conditioning polymers are used to enhance the<br />

conditioning properties, especially to mitigate the effects of extreme<br />

processing that are experienced during hair-straightening. Cationic<br />

polymeric conditioners can improve wet combability <strong>and</strong> ameliorate<br />

electrostatic charging of the hair (manifested by flyaway).<br />

Many cationic polymers have been developed for the purpose<br />

of conferring conditioning properties on hair. In fact, there are<br />

106


Chapter 3<br />

now more than one hundred polyquaternium ingredients listed in<br />

the INCI dictionary, <strong>and</strong> this list is still exp<strong>and</strong>ing. The primary<br />

purpose of polyquaternium polymers is to confer good conditioning<br />

benefits. A non-exhaustive list of conditioning polymers is shown in<br />

Table 1.<br />

Table 1. Examples of cationic conditioning polymers<br />

Chitosan<br />

Cocodimonium hydroxypropyl hydrolyzed Collagen<br />

Cocodimonium hydroxypropyl hydrolyzed hair Keratin<br />

Cocodimonium hydroxypropyl hydrolyzed Keratin<br />

Cocodimonium hydroxypropyl hydrolyzed Wheat protein<br />

Cocodimonium hydroxypropyl Oxyethyl Cellulose<br />

Steardimonium hydroxyethyl Cellulose<br />

Stearyldimonium hydroxypropyl hydrolyzed Oxyethyl Cellulose<br />

Guar hydroxypropyltrimonium Chloride<br />

Starch hydroxypropyltrimonium Chloride<br />

Lauryldimonium hydroxypropyl hydrolyzed Collagen<br />

Lauryldimonium hydroxypropyl hydrolyzed Wheat protein<br />

Stearyldimonium hydroxypropyl hydrolyzed Wheat protein<br />

polyquaternium-4<br />

polyquaternium-10<br />

Cationic hydroxyethylcellulose<br />

polyquaternium-24<br />

hydrophobically modified cationic hydroxyethylcellulose<br />

poly(methacryloxyethyltrimethylammonium methosulfate)<br />

poly(N-methylvinylpyridinium chloride)<br />

Onamer M (polyquaternium-1), peI-1500 (poly(ethylenimine)<br />

polyquaternium-2<br />

polyquaternium-5-poly(acrylamide-methacryloxyethyltrimethylammonium<br />

ethosulfate)]<br />

polyquaternium-6 poly(dimethyldiallylammonium chloride)<br />

polyquaternium-7<br />

poly(acrylamide-co-dimethyldiallylammonium chloride)<br />

107


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

Table 1. Examples of Cationic Conditioning Polymers (Cont.)<br />

polyquaternium-8<br />

polyquaternium-11<br />

108<br />

[poly-(N-vinyl-2-pyrrolidone-methacryloxyethyltrimethylammonium ethosulfate)]<br />

polyquaternium-16 [Co(vinyl pyrrolidone-vinyl methylimidazolinium chloride)<br />

polyquaternium-17<br />

polyquaternium-18<br />

polyquaternium-22<br />

poly(sodium acrylate – dimethyldiallyl ammonium chloride)<br />

polyquaternium-27<br />

polyquaternium-28<br />

polyvinylpyrolidone-methacrylamidopropyltrimethylammonium chloride)<br />

polyquaternium-31<br />

poly(N,N-dimethylaminopropylacrylate-N-acrylamidine-acrylamideacrylamidine-acrylic<br />

acid-acrylonitrile) ethosulfate<br />

polyquaternium-39<br />

poly(dimethyldiallylammonium chloride–sodium acrylate–acrylamide)<br />

polyquaternium-43<br />

poly(acrylamide-acrylamidopropyltrimoniumchloride-2-acrylamidopropyl<br />

sulfonate-DMapa)<br />

polyquaternium-44<br />

poly (vinyl pyrrolidone--imidazolinium methosulfate)<br />

polyquaternium-46<br />

poly (vinylcaprolactam-vinylpyrrolidone-imidazolinium methosulfate)<br />

polyquaternium-47<br />

poly (acrylic acid-methacrylamidopropyltrimethyl ammonium chloride–methyl<br />

acrylate)<br />

polyquaternium-53<br />

polyquaternium-55<br />

poly(vinylpyrrolidone-dimethylaminopropylmethacrylamide-lauryldimethylpropy<br />

lmethacrylamido ammonium chloride)<br />

pVp/Dimethylaminoethyl Methacrylate Copolymer<br />

Vp/DMapa acrylate Copolymer<br />

pVp/Dimethylaminoethylmethacrylate polycarbamyl<br />

polyglycol ester


Table 1. Examples of Cationic Conditioning Polymers (Cont.)<br />

pVp/Dimethiconylacrylate/polycarbamyl polyglycol ester<br />

Quaternium-80 (Diquaternary polydimethylsiloxane)<br />

poly(vinylpyrrolidone--dimethylamidopropylmethacrylamide)<br />

Vp/Vinyl Caprolactam/DMapa acrylates Copolymer<br />

amodimethicone<br />

peG-7 amodimethicone<br />

trimethylsiloxyamodimethicone<br />

Ionenes<br />

poly(adipic acid-dimethylaminohydroxypropyldiethylenetriamine)<br />

poly (adipic acid-epoxypropyldiethylenetriamine) (Delsette 101)<br />

Silicone Quaternium-8<br />

Silicone Quaternium-12<br />

Chapter 3<br />

Polyampholytes have been commercially available as<br />

conditioning polymers for a considerable time. A prominent<br />

example is polyquaternium-39, which is a copolymer of<br />

diallyldimethylammonium chloride, acrylamide, <strong>and</strong> acrylic acid.<br />

When this is polymerized in a single batch process, the mismatch<br />

in reactivity ratios between these monomers results in a lack of<br />

compositional uniformity. An improved version of this type of<br />

terpolymer of diallyldimethylammonium chloride, acrylamide, <strong>and</strong><br />

acrylic acid has been made by a monomer feed method for better<br />

control of molecular weight <strong>and</strong> composition. 95<br />

Copolymers comprising a diallylamine (typically diallyldimethyl<br />

ammonium chloride) <strong>and</strong> vinyllactam monomers (typically<br />

polyvinylpyrrolidone) are useful film-formers that confer<br />

conditioning properties such as good wet <strong>and</strong> dry combability, feel,<br />

volume, <strong>and</strong> h<strong>and</strong>leability. 96<br />

Silicone <strong>Conditioner</strong>s<br />

Silicone quaternaries have long been known as hair conditioning<br />

compounds.<br />

109


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

A recent patent application from Evonik Goldschmidt is<br />

directed to silicone quats that confer conditioning with longer<br />

lasting conditioning through several shampoo cycles. The premise<br />

is that long-term substantivity to hair requires the conditioning<br />

agent to contain a string of cationic charges. This was achieved<br />

by Goldschmidt by polymerizing cationic monomers <strong>and</strong><br />

grafting them to silicone backbones. In general, water-soluble<br />

monomers polymerized in the presence of silicones yield a<br />

mixture of water-soluble polymers <strong>and</strong> unsubstituted silicones<br />

because the two ingredients are incompatible <strong>and</strong> attachment<br />

of the polymer chain to the silicone would require appropriate<br />

“coupling groups.” The Evonik researchers rose to the challenge<br />

by polymerizing the cationic monomers in the presence of<br />

silicone polyethers. The ether groups are compatible with the<br />

quat monomers, <strong>and</strong> they readily chain transfer to give graft<br />

copolymers. Once grafted, the copolymers are quaternized to<br />

confer permanent positive charges with enhanced substantivity<br />

to hair. The grafts are obtained by polymerizing the readily<br />

available monomers, dimethylaminoethylmethacrylate, or<br />

3-trimethylammoniopropyl-methacrylamide.<br />

Leave-on silicone conditioners specifically targeted to nonshampoo<br />

applications confer enhanced <strong>and</strong> relatively durable<br />

conditioning. These contain emulsified vinyl-terminated silicones<br />

applied in combination with a conventional cationic conditioner. A<br />

preferred product type is a mousse. These silicone block copolymers<br />

can achieve excellent conditioning at relatively high viscosities (100<br />

KPa/s-1).<br />

Improved conditioning that confers surprisingly reduced friction<br />

on hair can be achieved by including an aminosilicone in which the<br />

aminosilicone has a fairly large range of average particle sizes from<br />

about 5–50 microns. 97<br />

References<br />

1. C Tanford, The Hydrophobic Effect: Formation of Micelles <strong>and</strong> Biological Membranes,<br />

Ch 1, Wiley Interscience: New Jersey (1980).<br />

110


Chapter 3<br />

2. HS Frank <strong>and</strong> MW Evans, Free volume <strong>and</strong> entropy in condensed systems: iii. Entropy<br />

in binary liquid mixtures; partial molal entropy in dilute solutions; structure <strong>and</strong><br />

thermodynamics in aqueous electrolytes, J Chem Phys, 13, 507-533 (1945).<br />

3. JT Davies <strong>and</strong> EK Rideal, Interfacial Phenomena, Ch 1, Academic Press: Waltham, MA,<br />

(1961).<br />

4. J Traube, Ueber die Capillaritätsconstanten organischer Stoffe in wässerigen Lösungen,<br />

Justus Liebigs Annalen der Chemie, 265, 27-55 (1891).<br />

5. JW Gibbs, The Collected Works of J. Willard Gibbs, Longmans: New York (1928).<br />

6. I Langmuir, The constitution <strong>and</strong> fundamental properties of solids <strong>and</strong> liquids, II. Liquids,<br />

J Amer Chem Soc, 39, 1848-1906 (1917).<br />

7. KJ Mysels, Soap Films: Studies of their thinning, Pergamon Press: Oxford (1959).<br />

8. JW McBain, Colloidal electrolytes: Soap solutions <strong>and</strong> their constitution, J Amer Chem<br />

Soc, 42, 426-460 (1920).<br />

9. GS Hartley, Aqueous Solutions of Paraffin-chain Salts, Hermann & Cie: Paris (1936).<br />

10. C Tanford, The Hydrophobic Effect: Formation of Micelles <strong>and</strong> Biological Membranes,<br />

pp 85, Wiley Interscience: New Jersey (1980).<br />

11. K Shinoda, N Yamaguchi, A Carlsson, Physical Meaning of the Krafft Point: Observation<br />

of Melting Phenomenon of Hydrated Solid Surfactant at the Krafft Point, J Phys Chem,<br />

93, 7216-7218 (1989).<br />

12. C Tanford, The Hydrophobic Effect: Formation of Micelles <strong>and</strong> Biological Membranes,<br />

pp 57, Wiley Interscience: New Jersey (1980).<br />

13. JN Israelachvili, DJ Mitchell <strong>and</strong> BW Ninham, Theory of self-assembly of hydrocarbon<br />

amphiphiles into micelles <strong>and</strong> bilayers, Faraday Trans. 2, J Chem Soc, 72, 1525-1568<br />

(1976).<br />

14. MG Cacace, EM L<strong>and</strong>au, JJ Ramsden, The Hofmeister series: salt <strong>and</strong> solvent effects on<br />

interfacial phenomena, Quart Rev Biophys, 30, 241-177 (1997).<br />

15. S C<strong>and</strong>au, A Khatory, F Lequeux, F Kern, Rheological behaviour of wormlike micelles:<br />

Effect of salt content, Journal de Physique IV, 03, C1, 197-209 (1993).<br />

16. JF Berret, J Appell, G Porte; Linear rheology of entangled wormlike micelles, Langmuir,<br />

9 (11), 2851–2854 (1993).<br />

17. E Buhler, JP Munch <strong>and</strong> SJ C<strong>and</strong>au, Dynamical properties of wormlike micelles: A light<br />

scattering study, J Phys II France, 5, 765-787 (1995).<br />

18. H Rehage <strong>and</strong> H Hoffmann, Rheological properties of viscoelastic surfactant systems,<br />

J Phys Chem, 92, 4712–4719 (1988).<br />

19. H Hoffmann, Viscoelastic surfactant solutions, in: Structure <strong>and</strong> Flow in Surfactant<br />

Solutions, CA Herb <strong>and</strong> RK Prud’homme, eds, ACS Symposium Series 578, American<br />

Chemical Society (1994).<br />

20. PA Hassan, SJ C<strong>and</strong>au, F Kern, <strong>and</strong> C Manohar, Rheology of wormlike micelles with<br />

varying hydrophobicity of the counterion, Langmuir, 14, 6025–6029 (1998).<br />

21. F Lequeux <strong>and</strong> S J C<strong>and</strong>au, Dynamical properties of wormlike micelles, in: Structure<br />

<strong>and</strong> Flow in Surfactant Solutions, CA Herb <strong>and</strong> RK Prud’homme, eds, ACS Symposium<br />

Series 578, American Chemical Society (1994).<br />

22. JF Berret, Transient rheology of wormlike micelles, Langmuir, 13, 2227–2234 (1997).<br />

23. NA Spenley, ME Cates, <strong>and</strong> TCB McLeish, Nonlinear rheology of wormlike micelles, Phys<br />

Rev Lett 71, 939–942 (1993).<br />

24. ME Cates, Theoretical modeling of viscoelastic phases, Structure <strong>and</strong> Flow in Surfactant<br />

Solutions, CA Herb <strong>and</strong> RK Prud’homme, eds, ACS Symposium Series 578, American<br />

Chemical Society (1994).<br />

25. NK Adam, J Soc Dyers Colour, 53, 122 (1937).<br />

26. E Kissa, Wetting <strong>and</strong> detergency, Pure & Appl. Chem, 53, 2255-2268 (1981).<br />

27. BJ Carroll, Equilibrium conformations of liquid drops on thin cylinders under forces of<br />

capillarity. A theory for the roll-up process.<br />

111


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

28. R Molina, F Comelles, MR Julia, P Erra, Chemical modifications of human hair studied by<br />

means of contact angle determination.<br />

29. V Dupres, D Langevin, P Guenon, A Checco, G Luengo, <strong>and</strong> F Leroy, Wetting <strong>and</strong><br />

electrical properties of the human hair surface: Delipidation observed at the nanoscale,<br />

J Colloid Interface Sci, 306, 34-40 (2007).<br />

30. CA Miller <strong>and</strong> KH Ramey, Solubilization-emulsification mechanisms of detergency,<br />

Colloids <strong>and</strong> Surfaces A, Physicochemical <strong>and</strong> Engineering Aspects. 74, 169-215 (1993).<br />

31. ASC Lawrence, The mechanism of detergence, Nature, 183, 1491 (1959).<br />

32. DG Stevenson, Surface Activity <strong>and</strong> Detergency, K Durham, ed., MacMillan: London<br />

(1961).<br />

33. CA Miller <strong>and</strong> KH Ramey, Solubilization-emulsification mechanisms of detergency,<br />

Colloids <strong>and</strong> Surfaces A, Physicochemical <strong>and</strong> Engineering Aspects. 74, 169-215 (1993).<br />

34. BJ Carroll, The kinetics of solubilization of nonpolar oils by nonionic surfactant solutions,<br />

J Colloid Interface Sci, 79, 126-135 (1981).<br />

35. JR Glenn et al, Porous, Dissolvable Solid Substrate <strong>and</strong> Surface Resident Coating<br />

Comprising Water Sensitive Actives, US Patent Application 20110195098, Aug 11, 2011;<br />

US Patent Application 20110189247, Aug 4, 2011; US Patent Application 20110189246,<br />

Aug 4, 2011; US Patent Application 20110182956, July 28, 2011.<br />

36. LS Tsaur et al, Personal Wash Cleanser Comprising Defined Alkanoyl Compounds,<br />

Defined Fatty Acyl Isethionate Surfactant Product <strong>and</strong> Skin or Hair Benefit Agent<br />

Delivered in Flocs Upon Dilution, US Patent Application 20110245125, Oct 6, 2011,<br />

assignee CONOPCO, INC., D/B/A UNILEVER.<br />

37. JJ Librizzi et al, Low-irritation compositions <strong>and</strong> methods of making the same, US Patent<br />

Application 20100311628, Dec 9, 2010, assignee Johnson & Johnson.<br />

38. ED Goddard, Polymer–surfactant interaction: Part II Polymer <strong>and</strong> surfactant of opposite<br />

charge, in: Interactions of Surfactants with Polymers <strong>and</strong> Proteins, ED Goddard <strong>and</strong><br />

KP Ananthapadmanabhan, eds, CRC Press (1993).<br />

39. ED Goddard, Polymer/surfactant interaction in applied systems, in: Principles of Polymer<br />

<strong>Science</strong> <strong>and</strong> Technology in Cosmetics <strong>and</strong> Personal Care, ED Goddard <strong>and</strong> JV Gruber,<br />

eds, Marcel Dekker (1999).<br />

40. P. Br<strong>and</strong>, et al, Cleansing formulations comprising non-cellulosic polysaccharide with<br />

mixed cationic substituents, US Patent 8,076,279, Dec 13, 2011, assigned to Hercules<br />

Inc.<br />

41. F. Utz, et al, Cationic cassia derivatives <strong>and</strong> applications therefore, US Patent 7,439,214,<br />

Oct 21, 2008, assigned to Lubrizol Advanced Materials, Inc.<br />

42. CA Lepilleur, Cationic cassia derivatives <strong>and</strong> applications therefore, US Patent 7,704,934,<br />

April 27, 2010; US Patent 7,923,422, April 12, 2011, assigned to Lubrizol Advanced<br />

Materials, Inc.<br />

43. E Baldaro, et al, Home And Personal Care Compositions, US Patent Application<br />

20110189248, Aug 4, 2011.<br />

44. MM Peffly, et al, Personal care compositions containing cationic synthetic copolymer<br />

<strong>and</strong> a detersive surfactant, US Patent Application 20070207109, Sep6, 2007, assigned to<br />

The Procter & Gamble Company.<br />

45. M. Pader, <strong>Shampoo</strong> compositions comprising saccharides, US Patent 4,364,837,<br />

Dec 21, 1982, assigned to Lever Brothers Company.<br />

46. RE Bolich, Jr., <strong>and</strong> TB Williams, <strong>Shampoo</strong> compositions containing nonvolatile silicone<br />

<strong>and</strong> xanthan gum, US Patent 4,788,006, Nov 29, 1988, assigned to the Procter &<br />

Gamble Company.<br />

47. J Parran, Detergent compositions containing particle deposition enhancing agents,<br />

US Patent 3,761,418, Sept 25, 1973, assigned to the Procter & Gamble Company.<br />

48. MJ Fair, et al, Bars comprising benefit agent <strong>and</strong> cationic polymer, US Patent 6,057,275,<br />

May 2, 2000, assigned to Unilever Home <strong>and</strong> Personal Care USA.<br />

112


Chapter 3<br />

49. MJ O’ Brien, Cosmetic compositions using polyether siloxane copolymer network<br />

compositions, US Patent Application, 2005/0197477, Sept 8, 2005, assigned to<br />

GEAM- Silicones.<br />

50. TN Biggs <strong>and</strong> GE LeGrow, Lightly cross-linked poly(n-alkylmethylsiloxanes) <strong>and</strong> methods<br />

of producing same, US Patent 5,493,041, Feb 20 1996, assigned to Dow Corning<br />

Corporation.<br />

51. J. A. Kilgour, et al, Branched organosilicone compound, US Patent Application<br />

2005/0165198, July 28, 2005, assigned to GE Plastics.<br />

52. SM Niemiec, et al, Novel detergent compositions with enhanced depositing, conditioning<br />

<strong>and</strong> softness capabilities, US Patent Application 2003/0176303, Sept 18, 2003.<br />

53. M Maubru, Cosmetic composition comprising at least one anionic surfactant, at least<br />

one cationic polymer <strong>and</strong> at least one amphiphilic, branched block acrylic copolymer<br />

<strong>and</strong> method for treating hair using such a composition, US Patent 7,498,022, March 3,<br />

2009, assigned to L’Oreal S.A.<br />

54. H Denzer, et al, Transparent aqueous compositions comprising hydrophobic silicone oils,<br />

US Patent 6,803,050, Oct 12, 2004, assigned to Kao Chemicals Europe S.L.<br />

55. MM Peffly <strong>and</strong> JE Hilvert, Conditioning shampoo compositions, US Patent Application<br />

20050158266, July 21, 2005, assigned to The Procter & Gamble Company.<br />

56. MM Peffly, et al, Clear personal care compositions containing a cationic conditioning<br />

polymer <strong>and</strong> an anionic surfactant, US Patent Application 20040234483, Nov 25, 2004,<br />

assigned to The Procter & Gamble Company.<br />

57. MM Peffly, et al, Clear personal care compositions containing a cationic conditioning<br />

polymer <strong>and</strong> an anionic surfactant, US Patent Application 20040234484, Nov 25, 2004,<br />

assigned to The Procter & Gamble Company.<br />

58. H Albrecht, et al, Hair shampoo containing pregelatinized, cross-linked starch derivatives,<br />

US Patent 7,279,449, Oct 9, 2007, assigned to Beiersdorf A.G.<br />

59. F Simonet <strong>and</strong> L Nicolas-Morgantini, Cosmetic composition of water-in-water emulsion<br />

type based on surfactants <strong>and</strong> cationic polymers, US Patent 20070237733, Oct 11, 2007.<br />

60. VM Coessens, K Matyjaszewski, Fundamentals of Atom Transfer Radical Polymerization,<br />

J Chem Ed, 87, 916-919 (2010).<br />

61. C Farcet, Gradient copolymer, composition including same <strong>and</strong> cosmetic make-up or<br />

care method; US Patent Application 20090047308, Feb 19, 2009, assigned to L’Oreal.<br />

62. N Cadena, et al, Method for depositing a polymer onto a surface by applying a<br />

composition onto said surface, USPatent 6,906,128, June 14, 2005, assigned to Rhodia<br />

Chemie.<br />

63. V Bergeron, et al, US Patent 7,335,700, Feb 26, 2008; US Patent 7,915,212, March 29,<br />

2011; <strong>and</strong> US Patent 7,939,601, May 10, 2011, assigned to Rhodia Inc.<br />

64. Q Stella, <strong>Shampoo</strong> containing a silicone in water emulsion, US Patent Application<br />

20030143177, July 31, 2003, assigned to the Procter & Gamble Company.<br />

65. DR Royce <strong>and</strong> RL Wells, <strong>Shampoo</strong> containing an alkyl ether, US Patent 7,087,221,<br />

Aug 8, 2006, assigned to the Procter & Gamble Company.<br />

66. DR Royce <strong>and</strong> RL Wells, <strong>Shampoo</strong> containing an ester oil, US Patent application<br />

20030138392, July 24. 2003, assigned to the Procter & Gamble Company.<br />

67. AJ O’Lenick, Jr, et al, Personal care products based upon surfactants based upon alkyl<br />

polyglucoside quaternary compounds, US Patent 6,881,710, April 19, 2005, assigned to<br />

Colonial Chemical Inc.<br />

68. MJ Cooke, et al, Conditioning shampoo comprising an aqeuous conditioning gel phase<br />

in the form of vesicles, US Patent Application 20110243870, Oct 6, 2011.<br />

69. S Midha, et al, <strong>Shampoo</strong> containing hollow particles, US Patent Application<br />

20030086896, May 8, 2003, assigned to The Procter & Gamble Company.<br />

70. S Midha, BD Hofrichter, Composition for improving hair volume, US Patent Application<br />

20030091521, May 15, 2003, assigned to The Procter & Gamble Company.<br />

113


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

71. RL Wells, ES Johnson, <strong>Shampoo</strong> containing a cationic guar derivative, US Patent<br />

6,930,078, Aug 16, 2005, assigned to The Procter & Gamble Company.<br />

72. SM Niemiec, et al, Detergent composition with enhanced depositing, conditioning <strong>and</strong><br />

softness capabilities, US Patent 6,495,498, Dec 17, 2002; US Patent 6,858,202, Feb<br />

22, 2005; US Patent 6,908,889, June 21, 2005, assigned to The Procter & Gamble<br />

Company.<br />

73. MF Neibauer, et al, Personal care compositions comprising a non-binding thickener with<br />

a metal ion, US Patent Application 20070009472, Jan 11, 2007, assigned to The Procter<br />

& Gamble Company.<br />

74. IC Constantinides, et al, Personal care compositions comprising responsive particles,<br />

US Patent Application 20070196299, Aug 13, 2007, assigned to The Procter & Gamble<br />

Company.<br />

75. T Yeoh, et al, Conditioning shampoo compositions having improved silicone deposition,<br />

US Patent 6,200,554; March 13, 2001, assigned to The Procter & Gamble Company.<br />

76. DR Weimer, Liquid detergent compositions, US Patent 3,718,609, Feb 27, 1973, assigned<br />

to Continental Oil Company.<br />

77. Olson, <strong>Shampoo</strong> composition possessing separate lotion phase, US Patent 3,810,478,<br />

May 14, 1974, assigned to Colgate-Palmolive Company.<br />

78. JR Williams, et al Separating multiphase personal care composition in a transparent or<br />

translucent package, US Patent 6,429,177, Aug 6, 2002, assigned to Unilever Home &<br />

Personal Care USA, division of Conopco, Inc.<br />

79. C Littau, et al, Multiphase aqueous cleaning compositions, US Patent 7,012,049, March<br />

14, 2006, assigned to Clariant International Ltd.<br />

80. KS Wei, et al, US Patent Application 20050100570, May 12, 2005, assigned to The<br />

Procter & Gamble Company.<br />

81. JA Wagner, et al, US Patent Application 20050192188, Sept 1, 2005, assigned to The<br />

Procter & Gamble Company.<br />

82. LS Tsaur, Personal polymer liquid cleansers stabilized with starch structuring system,<br />

US Patent 6,903,057, June 7, 2005, assigned to Unilever Home & Personal Care USA,<br />

division of Conopco, Inc.<br />

83. SPJ Ugazio, Polymer carriers <strong>and</strong> process, US Patent Application 20050203215,<br />

Sept 15, 2005, assigned to Rohm & Haas Company.<br />

84. RL Wells <strong>and</strong> ES Johnson, <strong>Shampoo</strong> containing a cationic guar derivative, US Patent<br />

6,930,078, Aug 16, 2005, assigned to The Procter & Gamble Company.<br />

85. Q Qiu, et al, Ordered liquid crystal cleansing composition with C16-24 normal<br />

monoalkylsulfosuccinates <strong>and</strong> C16-24 normal alkyl carboxylic acids, US Patent<br />

Application 20050136026, June 23, 2005, assigned to Unilever Intellectual Property<br />

Group.<br />

86. JJ Guth, DL Spilatro, RJ Verdicchio, Detergent compositions, US Patent 4,435,300,<br />

March 6, 1984, assigned to Johnson & Johnson Baby Products Company.<br />

87. SK Pratley, Aqueous cleansing composition, US Patent 6,001,787, Dec 14, 1999,<br />

assigned to Helene Curtis, Inc.<br />

88. MR Sivik, et al, Compositions <strong>and</strong> method for using amine oxide monomeric unit<br />

containing suds enhancers, US Patent 6,900,172, May 31, 2005, assigned to<br />

the Procter & Gamble Company.<br />

89. RW Glenn Jr., et al, Personal care compositions, especially those personal care<br />

compositions in the form of an article that is a porous, dissolvable solid structure,<br />

US Patent Application 20090232873, Sept 17, 2009.<br />

90. DJ Evans, et al, Separation <strong>and</strong> analysis of the surface lipids of the wool fiber, Proc. 7th<br />

Int. Wool Text. Res. Conf., Tokyo, Japan, I, 135–142 (1985).<br />

91. PW Wertz <strong>and</strong> DT Downing, Integral lipids of human hair, Lipids, 23, 878–881 (1988).<br />

114


Chapter 3<br />

92. PW. Wertz <strong>and</strong> DT Dowing, Integral lipids of mammalian hair, Comp Biochem Physiol,<br />

92B, 759–761 (1989).<br />

93. U Kalkbrenner, et al, Studies on the composition of the wool cuticle, Proc. 8th Int.<br />

Wool Text. Res. Conf., Christchurch, New Zeal<strong>and</strong>, I, 398 (1990); CM Carr, et al, X-ray<br />

photoelectron spectroscopic study of the wool fiber surface, Textile Res J, 56, 457<br />

(1986); S Breakspear, et al, Effect of the covalently linked fatty acid 18-MEA on the<br />

nanotribology of hair’s outermost surface, J Struct Biol, 149, 235–242 (2005); CA Torre,<br />

et al, Nanotribological effects of silicone type, silicone deposition level, <strong>and</strong> surfactant<br />

type on human hair using atomic force microscopy, J Cosmet Sci, 57, 37–56 (2006);<br />

M Yasuda, J Hair Sci, 95, 7–12 (2004); ML Tate, et al, Quantification <strong>and</strong> prevention of hair<br />

damage, J Soc Cosmet Chem, 44, 347–371 (1993).<br />

94. H Tanamachi, et al, Deposition of 18-MEA onto alkaline-color-treated weathered hair to<br />

form a persistent hydrophobicity, J Cosmet Sci, 60, 31–44 (2009).<br />

95. JJ Sabelko, et al, Low molecular weight ampholytic polymers for personal care<br />

applications, US Patent Application 20070207106, Sept 6, 2007, assigned to the Nalco<br />

Company.<br />

96. L Chrisstoffels, et al, Polymers comprising diallylamines, US Patent Application<br />

20070191548, Aug 16, 2007.<br />

97. ES Baker, et al, Conditioning composition comprising aminosilicone, US Patent<br />

8,017,108, Sept 13, 2011, assigned to the Procter & Gamble Company.<br />

115


<strong>Shampoo</strong> <strong>and</strong> <strong>Conditioner</strong> <strong>Science</strong><br />

116

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!