Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Advances in Colloid and Interface Science 256 (2018) 242–255

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science

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

Historical perspective

Interactions between surfactants and the skin – Theory and practice


Artur Seweryn
Department of Chemistry, Kazimierz Pulaski University of Technology and Humanities, Chrobrego Street 27, Radom 26-600, Poland

a r t i c l e i n f o a b s t r a c t

Available online 13 April 2018 One of the primary causes of skin irritation is the use of body wash cosmetics and household chemicals, since they
are in direct contact with the skin, and they are widely available and frequently used. The main ingredients of
Keywords: products of this type are surfactants, which may have diverse effects on the skin. The skin irritation potential
Surfactants of surfactants is determined by their chemical and physical properties resulting from their structure, and specific
Skin irritation interactions with the skin. Surfactants are capable of interacting both with proteins and lipids in the stratum
Anti-irritants
corneum. By penetrating through this layer, surfactants are also able to affect living cells in deeper regions of
Cosmetics
Household products
the skin. Further skin penetration may result in damage to cell membranes and structural components of
keratinocytes, releasing proinflammatory mediators. By causing irreversible changes in cell structure, surfactants
can often lead to their death. The paper presents a critical review of literature on the effects of surfactants on the
skin. Aspects discussed in the paper include the skin irritation potential of surfactants, mechanisms underlying
interactions between compounds of this type and the skin which have been proposed over the years, and verified
methods of reducing the skin irritation potential of surfactant compounds. Basic research conducted in this field
over many years translate into practical applications of surfactants in the cosmetic and household chemical in-
dustries. This aspect is also emphasized in the present study.
© 2018 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
2. Physical and chemical characteristics of surfactants in their aqueous solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
2.1. Classification of surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
2.2. Surfactant behavior in solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
3. Interactions between surfactants and the skin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
3.1. Interactions with SC surface proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
3.2. Interactions with SC lipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
3.3. Interactions with living skin cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
4. Methods to reduce the skin irritation potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
4.1. Use of surfactants with reduced skin irritation effect as primary washing agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
4.2. Application of surfactant mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
4.3. Addition of polymers and biopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
4.4. Addition of refatting and hydrophobic substances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
Declarations of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

1. Introduction through direct contact, the condition of the skin also depends on phys-
ical, biological and environmental factors [1,2]. Some of them may act si-
Every day the skin is exposed to aggressive factors that can cause cu- multaneously, thus potentiating the effect produced by individual
taneous irritation. In addition to chemical compounds affecting the skin factors. Examples include household chemicals containing abrasive in-
gredients or cosmetic scrubs. Abrasive substances used in both types
of formulations may cause abrasions and red patches on the skin
E-mail address: a.seweryn@uthrad.pl. resulting from the mechanical impact of individual grains on the surface

https://doi.org/10.1016/j.cis.2018.04.002
0001-8686/© 2018 Elsevier B.V. All rights reserved.
A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255 243

of the epidermis. Damage to the contact surface between the product components. Surfactants have an ability to alter the pH of the skin,
and the outer skin layer promotes an intensified activity of chemical fac- which results in abnormalities related to the synthesis of enzymes
tors which, in products of this type, are surfactants [1–5]. that are responsible for the production of intercellular matrix lipids
Surfactants have a wide range of applications in many industry and components of the NMF in the stratum corneum. Studies have re-
branches. They are used in the formulation of a vast number of commer- ported an increase in transepidermal water loss from the skin [46–48].
cially available products such as cosmetics and household chemicals. On The literature widely describes possible interactions between sur-
account of their properties determined by the amphiphilic structure factants and the skin which can result in dermal irritation. Surfactants
(ability to reduce surface tension at the interface between phases: are analyzed in the aspect of their binding to the SC surface proteins.
liquid–liquid, liquid–solid and liquid–gas, fat emulsification capacity), The process can lead to protein denaturation, changes in the structure
surfactants are a base for body wash cosmetics and a number of cleaning of the intercellular cement through the incorporation of surfactants
agents (e.g. hand dishwashing liquids, liquid laundry detergents) into its liquid crystalline structure followed by the washing and solubi-
[5–12]. Consequently, surfactants belong to the substances most com- lization of lipids which lead to damage to the epidermal barrier and in-
monly having contact with the skin. A key role in understanding the teractions between surfactants and living cells due to their penetration
mechanisms which underpin the skin-irritating activity of surfactants into deeper skin layers [1–5,9,13–40,46–49].
is attributed to the structure and physiochemical properties of their
aqueous solutions, an important factor being the form in which surfac- 2. Physical and chemical characteristics of surfactants in their
tants occur in the solution [1–5,9,13–38]. aqueous solutions
Surfactants interact primarily with the outermost epidermal layer
called the stratum corneum (SC). The stratum corneum consists of Surfactants are chemical compounds which have an ability to accu-
10–15 layers of corneocytes embedded in a lipid matrix (intercellular mulate (become adsorbed) at the interface between phases and reduce
cement). The SC is a barrier which prevents the entry of environmental surface tension. The property is attributable primarily to their bipolar
pollutants and pathogens into the body. Furthermore, it prevents exces- structure. Surfactant molecules contain two distinct parts differing in
sive water loss from deeper skin layers into the environment. polarity. The hydrophilic part has affinity to water and polar solvents,
Corneocytes are formed in the process of disintegration of cells which and forms the so-called head group of the surfactant molecule. It is com-
build up deeper layers of the epidermis (keratinocytes). Following the posed of (basic and acidic) groups that are capable of electrolytic disso-
loss of the cell nucleus and cytoplasm, keratinocytes are transformed ciation, and nonionic groups. The acidic groups are usually carboxylic,
into dead structures with tough outer cellular shells [39]. sulphonic, sulphate and phosphate radicals. The hydrophilic group com-
The intercellular cement has a unique structure. It is composed of posed of basic fragments includes, for example, primary, secondary and
alternating lipids, which are transported from dying cells in deeper tertiary pyridine groups and primary, secondary and tertiary amine
layers of the skin, and a thin layer of water saturated with considerable group. Hydroxide, sugar and thiol radicals represent nonionic groups.
amounts of various chemical compounds. The structure exhibits a The hydrophobic part which shows affinity to non-polar solvents and
rhomboid and hexagonal pattern of the lamellar phase, which makes oils typically consists of hydrocarbon chains. They can be fragments of
it highly resistant to the effect of a range of chemical agents and repre- fatty acids or their derivatives, either branched or unbranched.
sents the main factor inhibiting the migration of water from the dermis Also known are compounds in which the non-polar fragment is
to the epidermal surface [13,40]. The intercellular cement in well- composed of an aromatic carbohydrate with an attached long alkyl
structured stratum corneum contains a number of compounds includ- chain [10,11,50–58].
ing approximately 50% of ceramides, 25% of cholesterol, 15% of free
fatty acids and 10% of cholesterol esters [40–44]. 2.1. Classification of surfactants
Important constituents of the SC are substances present in
korneocytes forming the so-called natural moisturizing factor (NMF). The most commonly recognized classification of surfactants is based
They are hygroscopic compounds resulting from enzymatic transforma- on their ability to dissociate in water. Surfactants which undergo elec-
tions of a specific protein called filaggrin. The NMF consists primarily of trolytic dissociation in the aqueous environment are classified as ionic.
free amino acids (about 40%), sodium salt of pyroglutamic acid (about Where a negatively charged ion (anion) exhibits surface activity after
12%), lactic acid salts (about 12%), urea (about 7%), inorganic salts the process of dissociation, the compound is classified as an anionic sur-
(about 18.5%) and small amounts of ammonia, creatinine, uric acid, factant. Where surface activity is demonstrated for a cation, the surfac-
citric acid salts, sugars and peptides [13,40,43,44]. tant is cationic in nature. There are also compounds which are capable of
A major function of the epidermis is the formation of the so-called dissociation in water, however the type of ion formed depends on
hydrolipidic film, a protective layer at the contact surface between the the pH of the environment in which they occur. The compounds are
SC and the environment. The system contains mainly sebum and seba- classified as amphoteric surface-active agents. Accordingly, cations are
ceous gland secretions, making the skin soft and forming antibacterial formed in the acidic and anions in the basic environment. The dominant
and antifungal layers on the skin surface. It is a mixture of epithelial forms in solutions with a neutral pH are zwitterions, i.e. molecules con-
lipids, free fatty acids (about 5%), glycerides (about 50%), waxes taining both a cation and an anion. Those surfactants which dissolve in
(about 20%), squalene (about 10%), cholesterol esters (about 4%), water thanks to the formation of hydrogen bonds between water mole-
cholesterol (about 1%), sterols (about 1%) and other substances (about cules and an atom having a free electron pair are categorized as nonionic
9%). On the skin surface the mixture combines with a certain amount surfactants, as they fail to undergo electrolytic dissociation [11,50–58].
of water and sweat constituents, creating a W/O type emulsion. The General classification of surfactants is shown in Table 1.
formation of the emulsion is possible owing to the presence of sterols
which serve as W/O emulsifiers. The aqueous emulsion phase contains 2.2. Surfactant behavior in solution
dissolved substances which are secreted with sweat, most of which
have hygroscopic properties. They attract water from the environment From the physicochemical viewpoint, most household chemicals
and hinder its evaporation from the stratum corneum [45]. and body wash cosmetics can be regarded as aqueous solutions of sur-
An important role for preserving the structure and function of the factants enriched with various substances in order to achieve a desirable
SC is played by an appropriate pH gradient. It is recognized that the fragrance or viscosity, skin conditioning effect or product preservation
pH value on the skin surface should vary between 5.4 and 5.9. It is reg- [58–60]. In aqueous solutions surfactants undergo the process of self-
ulated among others by free fatty acids, metabolites of microorganisms organization which is a consequence of their amphiphilic nature.
residing on the skin, products secreted by endocrine glands and NMF Owing to their bipolar structure and the presence of two fragments of
244 A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255

Table 1
General classification of surfactants.

Surfactant Example
type

Anionic Sulphates (alkyl sulphates, alkyl ether sulphates), sulphonates (alkylbenzene sulphonates, α-olefin sulphonates, paraffin sulphonates, sulphonated fatty acids,
sulphonated methyl esters, sulphosuccinates), carboxylate (soap, ether carboxylates, acyl sarcosinates, taurates, isethionates), phosphate esters, acylated
aminoacids (N-acyl L-glutamates, N-acyl glycinates)
Cationic Long-chain amines and their salts, acylated diamines and polyamines and their salts, quaternary ammonium salts, polyoxyethylenated long-chain amines,
quaternized POE long-chain amines
Amphoteric Amidoamines, amidobetaines, sulfobetaines, sultaines, N-alkylbetaines, imidazoline carboxylates
Non-ionic Fatty alcohol ethoxylates, alkyl phenol ethoxylates, polyoxethylene esters of fatty acids, methyl ester ethoxylates, polyalkylene oxide block co-polymers,
amine ethoxylates, fatty alkanolamides, amine oxides, glycol esters, glycerol esters, polyglycerol esters, polyoxyalkylene polyol esters, alkyl poly glucosides,
polyoxyethylenated silicones

different polarities in their molecules, surfactants exhibit surface activ- The shape and size of arising micelles are determined by the type,
ity. In water they migrate and become adsorbed at the interface be- structure and concentration of their constituent surfactants. The shape
tween phases (e.g. water – air), where numerous active centres make of surfactant molecules is most commonly described by the packing fac-
the adsorption of surface active agent molecules possible. As the surfac- tor V/l ∙ S, where l indicates molecular length, V denotes the volume of a
tant concentration in the solution increases, and all the active centres single surfactant molecule and S – the surface area occupied by a single
become occupied, the concentration of compound molecules begins to molecule on the micellar surface. If only the molecular geometry is
rise in the volume phase of the solution, where they are present in the taken into account, the micellar shape in an aqueous solution can be
form of individual molecules called monomers. Surfactant monomers linked to the packing factor, as shown in Table 2 [6,58,63–65].
present in solutions are in constant motion. There is an ongoing Micelles formed in aqueous surfactant solutions after the CMC is
exchange between monomers in the volume phase and monomers exceeded are usually relatively small and spherical in shape or flattened.
adsorbed on the surface. The system tends to minimize interactions A characteristic feature of ionic surfactants is that all micelles formed in
which have an adverse effect on the system, occurring between water aqueous solutions are uniformly charged, which results in the develop-
molecules and hydrophobic groups of monomers present in the volume ment of mutual electrostatic repulsion forces attributable to the so-
phase. After a specific concentration, unique to each surfactant type and called zeta potential. The process prevents micellar aggregates from get-
referred to as the critical micelle concentration (CMC), is reached, ag- ting together at close distances, and stops them from fusing into large
gregates called micelles are formed in the solution [50–58]. Surfactant micelles. Repulsive forces exist not only between different micelles
behavior in solution is shown Fig. 1. but also between hydrophilic surfactant groups within a single aggre-
Micelle formation in the volume phase is determined by the so- gate. The interaction prevents the entry of a large number of its mole-
called hydrophobic effect which is associated with interactions taking cules into a single micelle (hence the value of the aggregation number
place between hydrophobic groups of surface active agents and water is constant) and is responsible for the continuous disintegration of
molecules. Since the system strives to reduce the surface of contact be- formed aggregates. The state of equilibrium enables only the exchange
tween solvent molecules and surfactant hydrocarbon chains, micelles of surfactant molecules between different micelles as well as micelles
develop in the solution [58,59]. Micellar aggregates consist of surfactant and monomers present in the volume phase. As soon as a micelle disin-
molecules which are oriented with their hydrophilic parts towards tegrates, a new aggregate is formed immediately in its place. In nonionic
solvent molecules, protecting the hydrophobic aggregate core from surfactants the repulsion of arising micelles is attributed to the spatial
contact with water. The system thus achieved is thermodynamically barrier that prevents aggregates from moving close to each other and
favourable. The quantity describing the mean number of surfactant mol- is referred to as the steric effect [61]. Since household chemicals and
ecules which can form a micelle is the so-called aggregation number. body wash cosmetics are systems comprising solutions of various sur-
Every surfactant type has its characteristic aggregation number, usually factant types, mixed micelles are formed, larger in size and higher in sta-
ranging from 50 and approximately 200. Micelles are thermodynami- bility [6,63–65]. The repulsion of such aggregates is attributed both to
cally unstable and, as a result, aggregates constantly disintegrate in the presence of the zeta potential and steric effects [7,8,49–58,62].
the solution and there is an ongoing exchange of individual molecules After the surfactant concentration in the solution rises above the
from the solution with aggregate-forming monomers. After the CMC is critical micelle concentration, the charges of individual aggregates grad-
exceeded, a state of equilibrium is achieved in the solution between ually change. Mutual repulsion forces between differently charged mi-
molecules of surface active agents in the form of monomers and mole- celles begin to disappear. Micelles can combine, which entails an
cules in micellar aggregates [50–58]. increase in the aggregation number and a marked enlargement in

Fig. 1. Surfactant behavior in solution. (a) Surfactant molecules in aqueous solution concentrate at the air–water interface with the hydrophobic part oriented towards the air side.
(b) When the concentration of surfactant increases, the interface saturates of surfactant molecules that penetrate into the solution. (c) To minimize their interaction with water, the
hydrophobic parts of the surfactants interact together and form micelles in solution.
A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255 245

Table 2 of skin irritations is associated chiefly with the presence of individual


Relationship between micellar shape and packing factor. surfactant molecules (monomers) in the aqueous solution. Rhein et al.
V/ls Surfactant molecular structure Micellar shape have shown that the skin irritation potential depends mainly on the du-
b0–1/3 Simple surfactants with a single Spherical or oval micelle
ration of skin contact with the surfactant before the CMC is achieved,
chain and appropriately large and the surfactant type and concentration. In their studies, the authors
polar group have found that after the CMC is exceeded, the severity of the irritant ef-
fect is lower. They have also argued that monomers are the main factor
1/3–1/2 Simple surfactants with a smaller Cylindrical micelles
determining the influence of surfactants on the stratum corneum's ker-
polar group or ionic surfactants
in the presence of electrolyte atin protein [66,67].
Monomers adsorbed on the skin surface may interact with the SC
1/2–1.0 Double-chain surfactants with Vesicles, bimolecular structures keratin protein, causing denaturation of its α-helix structure. Damage
a large polar group and flexible
to the secondary and tertiary structure makes it possible to expose
chains
sites where water molecules can bind. Swelling of the stratum corneum
is observed. The degradation of the protein structure facilitates the
washing of proteins from the skin and their solubilization in the solu-
1.0 Double-chain surfactants with a Planar (flat) or stretched micelles tion. As a result, it becomes easier for amphiphilic compounds, other
small polar group and stiff chains chemical compounds and pathogens to penetrate to deeper SC layers.
Reaching more deeply situated living cells, they are a stimulus inducing
an immune response, which is manifested by the development of topi-
cal red patches on the skin or itching. What is more, swollen keratin pro-
N1.0 Double-chain surfactants with a Reverse micelles teins after the evaporation of excess water have a lesser water binding
small polar group and large
non-polar group
capacity. As a result, the skin moisture level is lower, and the skin be-
comes less flexible [13,36,66–69].
The impact on proteins is attributed primarily to ionic surfactants
which are able to bind to them via relatively strong electrostatic interac-
tions [17,70–78]. Research on the impact of surfactants on the skin is fo-
cused above all on anionic surface active agents, with sodium dodecyl
their size. Since the system strives to achieve the smallest possible area sulphate (SDS) being most commonly used as a model compound. In
of aggregates, at a high concentration of surfactants it undergoes self- nonionic surfactants interactions with proteins are based on the forma-
reorganization. Large spherical micelles are transformed into cylindrical tion of relatively weak hydrogen and van der Waals bonds, which is
micelles and, as the next stage, into lamellar phases which are character- why their ability to cause protein denaturation and trigger skin irrita-
ized by a smaller area than large spherical micelles. A similar effect can tion through this mechanism is limited [13–17,19,29,36,66–68,79–85].
be achieved by introducing a salt (e.g. NaCl), used as a viscosity modifier, In recent years, the effect of surfactants on the skin has been evalu-
into the system. The addition of sodium chloride, which is most com- ated, both in scientific studies and practical applications, using the
monly used for this purpose, not only lowers the CMC but also, by reduc- zein protein [29,30,86–89] and bovine serum albumin [31,32,73,90,91]
ing the zeta potential in ionic surfactants and causing dehydration of which are structurally similar to keratin, the epidermis-building
polyoxyethylene chains in nonionic surfactants, it contributes to a protein. Measurements of the zein number involve determining the
change in the shape of micelles forming in the solution (spherical quantity of the water-insoluble zein protein solubilized by surfactants
micelles N cylindrical micelles N lamellar phases) [52–62]. The phenom- in the aqueous solution and previously partially denatured as a conse-
enon is used in production processes in order to achieve an appropriate quence of contact with an amphiphilic compound. The result is
viscosity level in cosmetic formulations [63–65]. The formation of cylin- expressed as the mass of free nitrogen derived from protein in the solu-
drical micelles in the system leads to an increase in viscosity [53–58]. tion and presented as a measure of the irritant potential. The higher the
Viscosity changes in aqueous surfactant solutions resulting from test result, the higher the skin irritation potential of the test sample
changes in micellar shape and size is shown schematically in Fig. 2. [29,30,63–65]. The interaction between bovine serum albumin and
surfactants has found practical applications in the evaluation of the
3. Interactions between surfactants and the skin skin irritation potential of cosmetic materials, finished cosmetics and
household chemicals. In the bovine serum albumin test, the structure
3.1. Interactions with SC surface proteins of albumin changes as a result of denaturation of the protein caused
by exposure to surfactants, ultimately leading to changes in the pH of
Surfactants rank among the most important factors with a potential the solution. The introduction of anionic surfactants into the albumin
to cause skin irritation. The literature data indicate that the occurrence solution causes surfactant binding to the cationic groups of the albumin

Fig. 2. Viscosity changes in aqueous surfactant solutions resulting from changes in micellar shape and size.
246 A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255

protein. In order to neutralize the negative protein charge resulting that an increase in the length of the sodium alkyl sulphate chain is ac-
from the predominance of anionic groups in its molecule, the adsorp- companied by an increase in the lipophilic properties of the compound,
tion of protons from the solvent takes place, causing a rise in the pH of which may be initially desirable in interactions with the stratum
the solution. The greater the increase in the pH level of the solution, corneum. However, as soon as the chain reaches C12 in length, any fur-
the stronger the skin irritation potential associated with the solution ther elongation causes the optimum lipophilicity value in terms of affin-
[31,32,91]. ity to the stratum corneum to be exceeded. A further increase in
According to the commonly recognized model of interactions be- molecule size contributes to an increase in hydrophobicity and restricts
tween surfactants and the SC proteins, only monomers – owing to its penetration to protein structures. The observations have been cor-
their small size – are capable of penetrating into the structure of the roborated by studies conducted by Di Nardo et al. [95] to evaluate
outer skin layer and inducing structural changes in constituent proteins. changes in skin hydration, TEWL and skin colour in 24-hour epidermal
It was believed that as soon as the critical micelle concentration (CMC) patch tests after the application of solutions of sodium alkyl sulphates
was achieved and micelles were formed, the severity of the irritant ef- containing chains of different lengths (C8, C12 and C14) at a concentra-
fect was reduced. The assumption was that micelles, being aggregates, tion of 0.2%. The highest increase in TEWL and the greatest decrease in
were too large to have an ability to deposit on or penetrate into the epi- the skin moisture level and in the severity of skin erythema were ob-
dermal surface. Furthermore, the potential to cause skin irritations was served for SDS. The effect of the length of the lipophilic chain (C8, C10,
also correlated with the value of the critical micelle concentration. In C12, C14, C16) on the ability to cause stratum corneum swelling has
surfactants with higher CMC values the concentration of free monomers been studied by Blake-Haskins et al. [33]. Their analysis of aqueous sur-
is higher, so the skin irritant effect increases as well [13–16,67–69,81]. factant solutions showed the maximum intensity of swelling for the ho-
The role of monomers as the only factor determining the skin irritant mologue with a chain length of C12. Similar findings have been reported
effect of surfactants in solutions has been verified by Walters et al. [82]. by Robbins and Fernee [96], demonstrating that among alkyl sulphates
The authors have found that the application of ready-made body wash the most severe adverse effect of surfactant homologues is associated
products in which the CMC is markedly exceeded also causes an in- with the alkyl chain containing 12 carbon atoms. In addition to the
crease in skin irritation, and the irritation potential rises as a function length of the CHO chain, the skin irritation potential of surfactants is
of surfactant concentration in the solution. Furthermore, they have also linked to the size of the polar part of the molecule. Based on
shown that both free monomers and micelles are capable of causing oxyethylates of sodium lauryl sulphate, Rhein et al. [67] have evaluated
protein denaturation. They claim that micelles, being unstable struc- how the degree of ethoxylation (1, 2, 3, 6 or 12 mol of ethylene oxide)
tures, may disintegrate after skin contact and affect its structure in the influences the stratum corneum swelling capacity. In their studies,
form of monomers [82,83]. In contrast, Rehin and Simion [66,67] pro- they have found that the ability to induce SC swelling decreases in pro-
pose that an increase in skin irritation activity noted after the CMC is portion to increasing degrees of ethoxylation. The correlation, however,
achieved can be attributed to small-sized submicelles emerging in the disappears after the ethoxylation degree exceeds 6.
solution or monomers released during micelle breakdown. Similar find- Similar studies have been conducted with Ammonium Laureth Sul-
ings have been reported by Moore et al. [81]. Their studies examining phate (ALES) homologues ethoxylated with 0, 3, 6 and 9 mol of ethylene
aqueous solutions of SDS at various concentrations have shown that oxide [33]. The least prominent swelling effect has been observed for
SDS is able to penetrate to the epidermis in monomeric and micellar the analogue with the greatest degree of ethoxylation (9OE). The
forms. The above findings have been confirmed by Cohen et al. [92] in highest skin irritation potential has been shown for ammonium lauryl
their most recent studies evaluating zein denaturation and solubiliza- sulphate (ALS), and has been found to decrease in line with increasing
tion capacity of sodium alkyl benzene sulphonate solutions. Exploring degrees of ethoxylation. Based on the studies presented above, it can
the correlation between zein and surfactant concentrations, the authors be concluded that an increase in the degree of ethoxylation is accompa-
have shown that the protein solubilization capacity increases over the nied by a decrease in the capacity of lauryl sulphates to induce SC swell-
CMC. Based on their findings, they confirm that surfactants in micellar ing and a reduction in their skin irritation potential. Such an effect can
forms are able to affect the epidermal protein structures. be determined by a rise in the degree of ethoxylation which has an im-
Some studies have shown that a significantly higher skin irritation pact on increasing the molecule size. Larger molecules exhibit a lower
potential is associated with those surfactants which form micelles of rel- ability to penetrate the skin's surface [33,67].
atively small size, with a correlation found between the study results Studies have also been conducted to analyze solutions of different
and the surfactant structure [13,14,19,33,66,81]. For example, the skin lauryl sulphate salts (Na, Mg, TEA) with a focus on their ability to
irritant effect induced by sodium dodecyl sulphate (SDS) increases cause stratum corneum swelling. The most pronounced swelling effect
quite considerably even after the CMC is exceeded. However, the in- has been shown for the sodium salt and the least pronounced for mag-
crease is much lower for ethoxylated sodium dodecyl sulphate. In the nesium alkyl sulphate [67]. Similar findings have been presented by
case of SDS the increase in the skin irritation potential is due to the Cohen et al. [97], in their studies evaluating the effect of counterion in
small micelle size, since micelles are smaller than the orifices of hair fol- alkyl sulphates on the skin irritation potential by determining the zein
licles. In the opinion of the authors, the latter represent another route number and correlating their results with measured CMC values and
for the entry of SDS into the structures of the stratum corneum. In the ionic radius size. The authors analyzed the effect of inorganic
other surfactants the size of micellar aggregates exceeds the diameter metal ions including Li+, Na+, K+, Mg2+, and NH4+, and organic ions
of hair follicle orifices, making the process of penetration more difficult. including MEA+, DEA+ and TEA+. Among the compounds studied, the
The findings thus show that the formation of micelles in the solution lowest zein number values were shown for the magnesium salt. Other
does not aid in achieving total inhibition of the skin irritant effect salts with inorganic monovalent ions, however, exhibited no significant
[13,14,66,67,81,93,94]. differences in value. Organic salts were found to produce significantly
The skin irritation potential is significantly dependent on the chem- higher zein numbers than inorganic salts. According to the researchers,
ical structure of surfactants. Wilhelm et al. [68] have studied changes in the observation can be attributed to stronger binding of these ions in
transepidermal water loss (TEWL) in 24-hour epidermal patch tests micelles. Both the specific ability to cause the SC swelling and the zein
performed on skin subjected to sodium alkyl sulphates with alkyl chains number results were linked by the authors to the CMC values and ion
of different lengths (C8, C10, C12, C14, C16). They have found that the size. The highest tolerability of magnesium salts by the skin may be
value of TEWL rises with increasing chain length, however only up to due to the fact that their aqueous solutions have lower CMC values
C12, with the highest value recorded for SLS. Subsequent homologues and, consequently, they are characterized by lower monomer concen-
with longer lipophilic chains (C14 and C16) cause a smaller increase trations in the system. According to another hypothesis, ion size is the
in this parameter. Based on the findings, the authors have concluded factor which determines the extent of interactions with the stratum
A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255 247

corneum. In magnesium ions an excessive volume of ionic aggregates and micelles formed in the solution. A number of experimental studies
(magnesium having the largest ionic radius of all studied) prevents have shown that exposure to a surfactant solution in which the CMC
easy diffusion in the stratum corneum [67,97–99]. Similar results have is exceeded leads to intemicellar solubilization of lipids and their
been obtained by Seweryn et al. [100] in applied studies evaluating washing from the stratum corneum. Possible consequences include
the skin irritation potential of hand dishwashing liquids in the coacer- delipidation of the stratum corneum, impairment of its barrier functions
vate form obtained through the addition of various inorganic salts (of and increase in the TEWL level. Some studies also postulate that mono-
mono- and divalent metals). The evaluation of the irritant potential mers, which are small-sized molecules, may become adsorbed and em-
based on determining the zein number produced significantly lower bedded into the liquid crystalline structure of the stratum corneum,
values of the parameter studied in the products obtained by adding disturbing its spatial structure and increasing permeability. In extreme
magnesium chloride and calcium chloride. cases, following a prolonged contact between the SC and a surfactant so-
A specific class of amphiphilic substances comprises gemini surfac- lution, the intercellular cement can become liquefied and the epidermal
tants. Their chemical structure translates into a range of more beneficial lipids can be more prone to solubilization. As the permeability of the
properties compared to their more conventional counterparts. Owing to epidermis increases, monomers are able to migrate into deeper-lying
growing interest in surfactants of this type in the context of their prac- regions of the stratum corneum. As a result, the skin irritation effect in-
tical applications in the cosmetic industry, a number of studies focusing tensifies. In addition, water molecules are able to migrate to the surface
on their interactions with the skin have been published, demonstrating of the epidermis, resulting in greater evaporation and dryness of the
that compounds of this type induce a negligible skin irritation effect stratum corneum. The skin becomes dry and loses its elasticity
[101–112]. As Tsubone et al. [111] argue, since gemini surfactants [13–15,36,40,128,129].
have markedly lower CMC values than conventional surfactants, it fol- Monomers present in the structure of the intercellular cement may
lows that they are also associated with a lower skin irritation potential. damage enzymes that produce lipids forming the matrix. Changes to
In vivo tests performed for synthesized gemini-type anionic surfactants the composition and content of different lipids are observed. Fulmer
have revealed low levels of skin irritation. In addition, the authors point et al. [130] have studied the effect of SDS on the quantitative and qual-
to a correlation found between their results and CMC values calculated itative composition of lipids in the stratum corneum, noting that SDS
for the compounds studied. Diz et al. [112] have evaluated cationic contributes to a decrease in the content of long-chain free fatty acids
gemini-type surfactants in Draize tests, showing antibacterial activity and cholesterol, without affecting the total content of ceramides and
as well as the absence of any skin irritation effect at concentrations non-polar lipids. Similar findings have been presented by Froebe et al.
below 0.5%. [129] based on in vivo studies evaluating the correlation between the ef-
The developments of biotechnology as a science and the search for fect of sodium lauryl sulphate and sodium alkyl benzene sulphonate on
raw materials of natural origin have sparked a rapid expansion of the SC proteins and lipids. The authors found that the lipids were
research into the development of surfactants of natural origin. One of washed away only at or above the concentration at which micelles
the most promising areas within this field of study relates to formed in the system. Measurements of the SC composition after 20-
biosurfactants. They are compounds of biological origin, usually pro- minute exposure to 2% surfactant solutions revealed a 7% decrease in
duced by microorganisms such as bacteria, yeasts and moulds. the total lipid content. An assessment of different lipid fractions demon-
Biosurfactants are not only effective at reducing the surface or interfa- strated the elimination of free fatty acids (8%), cholesterol (8–15%), and
cial tension, which is a characteristic feature of all surfactants, but they cholesterol and squalene esters (under 1%), however ceramides were
are also non-toxic and highly biodegradable. Therefore, as opposed to not shown to be washed away. According to the researchers the lack
synthetic surfactants, they do not pose a threat to living organisms of interactions with the ceramide fraction is due to ceramides having
[113,114]. The beneficial properties of biosurfactants, coupled with the two carbohydrate chains, which reduces their potential for solubiliza-
current development trends in cosmetics and household chemicals, tion in micelles. The authors point out that even though the elimination
have brought about studies exploring the possibilities for using surfac- of short-chain and unsaturated fatty acids is desirable in day-to-day hy-
tants of this type in production. An inherent element of this line of re- gienic activities, the loss of stearic acid and other long-chain fatty acids
search involves tests assessing the effect of biosurfactants on the skin. may adversely affect the barrier function of the stratum corneum
Published research results show that biosurfactants are characterized [36,40,129,130].
by limited skin irritation and toxicity, as compared to synthetic surfac- Based on the findings presented above, it may be concluded that
tants [113–121]. In addition, some of them have antioxidant properties fatty acids are the most prone to potential interactions with surfactant
and improve the bacterial microflora of the skin [113–115,122–127]. solutions. A precondition for the preservation of the bi-layer liquid
crystalline structure of the intercellular matrix in the SC is to ensure ap-
3.2. Interactions with SC lipids propriate preferential ratios between the ceramide, fatty acid and cho-
lesterol fractions. If one of the constituents is removed, the entire
Research investigating possible interactions between surfactants or system may become destabilized and defective. The removal of fatty
body wash products and the skin is largely focused on the evaluation acids from the remaining components may adversely affect the preser-
of skin irritations induced by interactions with proteins. Less attention vation of the desirable liquid structure, which may result in excessive
has been devoted to potential interactions with the liquid crystalline stiffness of the SC and reduced ability for normal tension relaxation pro-
lipid structure of the intercellular cement within the SC and its damage. cesses. Also, aside from an increase in TEWL, dry skin sensation, cracking
The first study in this area, by Imokawa et al., has investigated the im- and, in extreme cases, erythema, normal epidermal desquamation may
pact of surface active agents on the SC lipids in correlation with ob- become disturbed. The elimination of intercellular matrix lipids bound
served skin dryness [128]. covalently to corneocytes interferes with the perfect layered arrange-
It was previously believed that the interactions between surfactants ment of cells forming the epidermis. Corneocytes can get closer to one
and the SC lipids were limited exclusively to the processes of washing another and migrate between layers, aggregating into larger cell clus-
lipids forming the epidermal hydrolipid film. They were related to ters. The process of desquamation involves non-uniform shedding of
such aspects as daily hygiene, skin defatting, and elimination of sweat, skin scales [36,40,128–131].
sebum and contaminants. It is currently proposed that surfactants,
being surface active agents, are capable of interacting with lipids 3.3. Interactions with living skin cells
which form the SC intercellular cement [14–15,36,40,128,129].
The commonly recognized model of interactions takes into account The denaturation of proteins building the stratum corneum, and
both possible effects produced by surfactants in the form of monomers interactions between the intercellular substance and surfactants,
248 A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255

can cause damage to the skin's protective barrier. As a result of the pro- Allergies resulting from skin contact with surfactants are noted rela-
cesses, the compounds are able to migrate more easily into deeper epi- tively rarely, however there are literature reports describing skin intol-
dermal layers, where keratinocytes, i.e. living epidermal cells, are erance to a range of compounds including sodium lauryl sulphate
present. Surfactants are capable of binding to the keratinocytic cyto- [19,138,146], cocamidopropyl betaine [147], alkyl polyglycosides
plasm and causing its impairment. They can also damage proteins [148–150] or sarcosinates [151]. More common allergy-causing agents,
which build living cells. The processes may lead to permanent destruc- which occur both in cosmetics and household chemicals, are colourants,
tion of keratinocytes or even cell death. The capacity of surfactants to in- preservatives, plant extracts and fragrances [152–154].
duce permanent damage to living skin cells is referred to as skin toxicity
[132–137]. Furthermore, the presence of surfactants and their interac- 4. Methods to reduce the skin irritation potential
tions with proteins can cause damage to the structure of living epider-
mal cells, producing topical inflammatory reactions of the skin According to the literature, in order to reduce the irritation potential
[13–15,134,137,138]. This type of reaction to an irritant factor is re- of surfactants or products based on anionic surface active agents, at-
ferred to in the literature as irritant contact dermatitis (ICD), which be- tempts should be made to lower the concentration of monomers, in-
longs to the most common cutaneous diseases [134–142]. crease the size of micelles and stabilize them in the solution having
Depending on the type and concentration of surfactants, and the du- contact with the skin. There are various methods comprising the steps
ration of exposure to the irritating factor, the process involves presented above which make it possible to reduce the negative impact
keratinocyte damage. It affects primarily the cellular cytoplasm, causing of surfactants on the skin in body wash products. For example, anionic
the release of inflammatory mediators from cells, including various pro- surfactants with a proven mild effect on the skin, mixtures of different
inflammatory cytokines such as interleukin 1 (IL-1), IL-6, IL-8 and tu- surfactant types, an addition of polymers or refatting and hydrophobic
mour necrosis factor (TNF). The presence of cytokines results in the substances can be used as the primary washing agent in the formula-
dilation of blood vessels and in the influx of immune cells which induce tions [1,13–16,155–159].
the inflammatory process. They are mostly neutrophilic granulocytes,
and T and B lymphocytes. Concurrently, in order to reduce excessive in- 4.1. Use of surfactants with reduced skin irritation effect as primary
flammatory response, mechanisms are set in motion that involve washing agents
keratinocytes producing cytokines with antiinflammatory properties.
ICD-associated inflammatory processes occur exclusively within the A reduced skin irritation effect of cosmetic products can be achieved
site of contact with the irritating substance. They result in topical skin by an appropriate selection of product components. A proper choice of
lesions presenting as red skin, desquamation and erythema. In some the type of primary surfactants seems to be a fundamental means to ob-
cases patients develop only subjective sensations including severe pru- tain desired functional parameters and reduce the irritant effect on the
ritus, pain, burning and feeling of discomfort. An important factor is that skin induced by products of this type. Both appropriate detergent and
ICD does not have immune aetiology, which means that the body pre- foaming properties which are so sought-after in cosmetics, and a mild
serves no information about the skin-penetrating compound in the effect on the skin produced by surfactants, are linked to their structure.
form of antigens [138–143]. As mentioned, surfactants which interact strongly with proteins in
The literature reports suggest that irritant contact dermatitis is one of the stratum corneum, causing their denaturation and swelling, can
the most common skin diseases. As Levin and Maibach [143] claim, aside be regarded as compounds with a high skin irritation potential
from the chemical structure of the triggering substance, the occurrence [1,14,15,17]. It is possible to determine the tendency of surfactants to ir-
of ICD is determined by a range of other factors such as age, sex, genetic ritate human skin by analyzing a range of aspects which are widely de-
background, environmental conditions and coexistence of other cutane- scribed in the literature, including a higher tendency of surfactants to
ous diseases such as atopic dermatitis. The disease must therefore be induce the swelling of keratin in the stratum corneum or model proteins
recognized as multifactorial. A greater incidence of disease symptoms (keratin and collagen) [96,160], a specific effect of surfactants on the de-
has been found in people who are exposed to cosmetics and cleaning naturation of the globular protein bovine serum albumin [31–35] or
agents on a prolonged basis, for example hairdressers and cleaners. their effectiveness in causing partial denaturation and solubilization of
Another effect of interactions between surfactants and living the water-insoluble zein protein [29,30,87–92,100].
epidermal cells which, however, occurs very rarely, is allergic contact The molecular structure of surfactants is essentially the most impor-
dermatitis (ACD) [139,140,144,145]. As opposed to ICD, ACD is an tant factor for the problem of surfactant-associated skin irritation effect
immunologically induced syndrome in which the immune system re- which is discussed in the present paper. Generally, ionic surfactants
sponds to a chemical substance referred to as an allergen. The mecha- have a more pronounced irritant effect on the skin, which is attributable
nism underlying contact allergy comprises two phases of body to the nature of their interactions with proteins via electrostatic and hy-
response to a skin-penetrating substance (induction and release drophobic bonds. However, unlike in the ionic type, the strength of in-
phases) that involve complex physiological and biological processes. teractions in nonionic surfactants is low on account of the absence of
In the first phase, the allergen is phagocytized by Langerhans cells, pre- electrostatic forces (resulting from the lack of charge). The binding of
senting it to B lymphocytes, which are transformed into immunoglobu- compounds of this type with proteins via hydrophobic interactions
lin proteins (antibodies), which are specific and sensitive to particular and hydrogen bonds does not strongly affect the protein structure
allergens. They are transported by the cardiovascular system all over [66–84]. Consequently, they are considered to be non-irritant on the
the body and presented to T lymphocytes – immune system cells skin. In general terms, the tendency for protein interactions follows
which retain information about the allergen. The process is symptom- this sequence:
less and lasts from a few to about a dozen days. The phase of release in- anionic = cationic N amphoteric N nonionic surfactants.
volves the recognition of T helper cells of the allergen by the immune Cationic surfactants are not used as primary washing agents in
cells, accompanied by an increase in the production of histamines and washing products. On account of the fact that the surface of the skin
inflammatory cytokines. A severe inflammatory response develops not and hair is negatively charged, they may bind relatively strongly to
only at the site of entry but also on the entire skin surface. The processes such surfaces. Positively charged hydrophilic heads of these surfactants
lead to cutaneous eruptions in the form of bullae, red skin, oedema, and bind to keratin, whereas hydrophobic tails are oriented towards the
the sensation of itching. Similarly to ICD, the development of ACD de- outside. Consequently, this surface has a more hydrophobic character.
pends on a number of factors such as age and sex, however the greatest As a result, the skin and hair conditioning effect is excessively increased,
role is attributed to genetic predisposition related to the mutation of ap- but also the attraction of hydrophobic dirt is enhanced, making it more
propriate genes [139,140,144,145]. difficult to remove such dirt and increasing the tendency for its
A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255 249

redeposition on the surface [161]. There is a lot of contradictory infor- agent concentration. Generally, binding isotherms have four character-
mation about the irritant effect of cationic surfactants. Some claim that istic regions including: specific binding (a), non-cooperative binding
they have a more severe irritant effect on the skin and hair than anionic (b), cooperative binding (c) and saturation (d).
surfactants [136,162–165] whereas in zein solubilization tests, corre- Specific binding is mainly electrostatic in nature [168]. Surfactant
lated with the irritant effect, the protein solubilization capacity of cat- headgroups bind to oppositely charged groups on the protein. A change
ionic surfactants is almost twice as low as that of anionic surfactants in pH entails a change in the net charge of the protein, leading to surfac-
[1,13,14]. It seems evident that strong interactions of cationic surfac- tant binding to the protein. If the pH is lowered, the binding isotherm of
tants with proteins and their lack of effective detergent activity deter- an anionic surfactant is shifted to lower surfactant concentrations [169].
mine the application of surfactants of this type exclusively as auxiliary In contrast, the binding isotherm for cationic surfactants is observed at a
conditioning agents in body wash cosmetics, lotions and hair condi- higher concentration of the amphiphilic compound [170]. Specific bind-
tioners. Strong protein denaturation ability exhibited by some cationic ing (region a) of surfactants to proteins depends on the headgroup of
surfactants translates into their antibacterial activity. the ionic surfactant and may be selective in nature. In their study,
Amphoteric surfactants are rarely used as the main ingredients of Reynolds et al. [171] report that the structure of bovine serum albumin
cosmetic products, and they usually serve as auxiliary agents in such (BSA) comprises binding sites which do not bind alkyl carboxylates at
formulations [10,11,21,58]. Depending on the pH level, they may be all or do so with a reduced affinity, whereas both alkyl sulphates and
positively or negatively charged, which affects the nature and strength sulphonates are strongly bound. However, for specific binding to
of interactions with proteins. Generally, compounds of this type have occur, the length of the hydrophobic part of the compound seems to
a relatively low irritant potential compared to cationic and anionic sur- be an important factor. Jones and Manley [169] report that at low con-
factants. In practical terms, in the context of interactions with the skin, centrations of n-dodecyl and n-decyl sulphates a specific binding to
compounds of this type are used in products in combination with an- the lysozyme is observed, however at higher concentrations of the com-
ionic surfactants. By producing a synergistic effect, especially in mildly pounds the binding is cooperative (region b). In contrast, sodium octyl
acidic environments where a zwitterionic surfactant may have a cat- sulphonate binds to protein exclusively via cooperative bonds. Similar
ionic charge, they contribute to reducing the irritant effect of the formu- findings were noted for the binding of n-alkyl trimethylammonium bro-
lation [1,14,166,167]. mides to BSA [172]. For decyltrimethyl ammonium bromide, no transi-
In view of the absence of interactions with proteins, nonionic surfac- tion from specific binding (region a) to cooperative binding was
tants are generally regarded as mild-acting [13–17,19,29,36,66–68, observed, unlike in compounds with a longer alkyl chain – dodecyl
79–84]. Based on that property, they are used in the formulation of and tetradecyltrimethyl-ammonium bromide. On the other hand, how-
body wash products intended for babies and children [82]. On the ever, in BSA specific binding is observed both for sodium dodecyl and
other hand, however, compounds of this type are good emulsifiers, decyl sulphates, and sodium octyl sulphonates [173]. The number of
which may produce an adverse effect on the lipid components of the specific bonds with albumin (BSA) rises along with the increasing
skin's protective barrier and enhance the process of washing away in- length of the hydrocarbon chain for the homologous series of n-alkyl
tercellular matrix lipids. Furthermore, alkyl polyglycosides, on account sulphates and sulphonates [171]. In addition, binding is initiated at a
of their high substantivity to the surface, when not completely removed lower surfactant concentration along with an increase in chain length
from the surface, may play a part in intensifying the process of emulsi- [174].
fication of hydrophobic substances during the supply of consecutive Cooperative binding (region c) means an increased affinity for binding
amounts of water during subsequent washing. to protein along with an increase in the concentration of free surfactant. A
The most widely employed surfactants in body wash cosmetics sharp rise in the isotherm is noted within a narrow concentration range.
and selected types of household chemicals are anionic surfactants. The formation of micelle-like structures of surfactants is observed on
For this reason, their potential to interact with protein structures of the protein, representing a cooperative process. The protein may unfold
the skin and trigger skin irritations represents the most significant in a specified region. Takeda et al. [175] studied changes in the
problem. structure of bovine serum albumin in solutions of sodium decyl sulphate
The general model of interactions between ionic surfactants and (SDeS), sodium dodecyl sulphate (SDS), decyltrimethylammonium
proteins can be based on a binding isotherm [86,168] (Fig. 3.) which bromide (DeTAB), dodecyltrimethylammonium bromide (DTAB),
shows the average number of surfactant molecules bound per protein tetradecyltrimethylammonium bromide (TTAB) and hexadecyltrimethyl-
molecule (ν) as a function of the logarithm of the free surface active ammonium bromide (HTAB). All these surfactants caused changes in the
helical structure proportion in the protein from 66% to approximately
50%. However, the shorter the hydrocarbon chain in the structure of the
surface active agent, the higher concentration is required to cause an
equivalent decrease in the proportion of this structure. Study results
also indicate that the longer the hydrocarbon chain in a compound, the
greater the drop in the proportion of the helical structure. A mild increase
in the protein β-structure was observed for all surfactants except for
DeTAB.
A number of proteins (globular, fibrous, membrane proteins) bind
identical amounts of SDS at the saturation state (region d), with approx-
imately 1.4 g of SDS per 1 g of protein. The saturation with anionic sur-
factants is pH-independent and seems to be controlled by cooperative
hydrophobic bonds. For cationic surfactants, a reduced intensity of sat-
uration binding is noted in comparison with anionic surfactants [176].
Interactions taking place between ionic surfactants and proteins in-
volve primarily the effect of hydrophobic parts on hydrophobic protein
fragments and electrostatic binding of hydrophilic, charged surfactant
heads to some charged protein parts. It is, therefore, recognized that a
limited impact on the protein structures – and hence (in the case of
Fig. 3. The isotherm of the binding of ionic surfactants to proteins. Own study based on skin proteins) a reduced irritant effect on the skin – occurs in ionic sur-
[168]. factants with a lower hydrocarbon chain number and length and a
250 A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255

reduced charge density in the hydrophilic part, for example through the micellar stability and leads to its relatively fast breakdown into mono-
incorporation of an additional hydrophilic part in the form of ethylene mers. Following the addition of cationic surfactants (B) to a solution of
oxide groups [33,67,86–99] Some research in this field indicates that anionic surfactants, the former become embedded into the micelles
surfactants exhibiting higher self-assembling ability have a weaker irri- formed in the solution, giving rise to mixed micelles. Aggregates of
tant activity [163,177,178]. Surfactant-induced protein denaturation this type, containing both surfactant types, are much larger in size.
can be characterized through cooperative surfactant adsorption The forces of mutual repulsion are reduced, which has an additional sta-
[29,30,179]. If the functional group of the ionic surfactant is large in bilizing effect on the micelles. After adding a nonionic compound (C) to
size, interactions between the hydrophobic part of its molecule a solution of anionic surfactants, large-sized mixed micelles are formed,
and the hydrophobic protein fragment are limited, and no adsorption as in the case of adding cationic surface active agents. Molecules of non-
on the protein is observed [86]. Furthermore, weak repulsion between ionic surfactants, which are relatively large, increase the distances be-
surfactants adsorbed on the protein induces minor changes in the tween hydrophilic heads of ionic surfactants, at the same time
protein secondary structure, either diminishing or eliminating coopera- weakening the forces of electrostatic repulsion. As a consequence, mi-
tive adsorption [87]. The above findings are consistent with the obser- celle stabilization occurs. A similar effect is attained by combining an an-
vations made by Ozawa et al. [157] in their study of five anionic ionic surfactant solution with another anionic surfactant (D). On
surfactants: sodium polyoxyethylene lauryl ether carboxylate, sodium account of a difference in the structure of the hydrophilic and hydro-
polyoxyethylene alkyl ether sulphate, sodium dodecyl sulphate, potas- phobic segments, the distance between the hydrophilic parts grows,
sium laurate, and N-cocoyl-L-glutamic acid monosodium salt. The au- which increases the stability of micelles. However, it needs to be
thors correlated their findings obtained for epidermal swelling and stressed that the level of stabilization thus achieved is markedly lower
protein denaturation as determined by the zein test with the results than that noted after the addition of nonionic and cationic surfactants.
or surface tension and stratum corneum adsorption. The study showed The stability of micelles in systems of this type is additionally affected
sodium polyoxyethylene lauryl ether carboxylate to have the lowest by the size difference of the hydrophobic part of anionic surfactants –
skin irritation effect of all the surfactants studied. The finding was attrib- the greater the difference in the length of the hydrocarbon chain, the
uted to the size of the ionic functional group and the ability of the com- higher the level of micellar stability. An addition of amphoteric surfac-
pound to form specific aggregates with a vesicular structure in aqueous tants (E) induces an increase in micellar size, however the stability of
solutions even below the CMC. Similar conclusions were presented by micelles depends on the pH of the system. In the acidic environment,
Zhou et al. [163] who studied cyclodextrin/cationic trimeric surfactant an amphoteric compound carries a positive charge and observed
complexes. The authors point to the possibility of large aggregate sys- changes are the same as after the addition of cationic surfactants (B).
tems of this type forming in aqueous solutions directly after the CAC is In the basic environment, the effect is similar to that occurring after
achieved, and a decrease in the concentration of free monomers. Strong the addition of an anionic surfactant (D) [14,52–60].
self-assembly reinforces the repulsion of complexes from the zein The theories outlined above are confirmed by a range of studies.
backbone. Furthermore, a reduced number of hydrophobic chains and Hall-Manning et al. analyzed the irritant effect induced by a mixture
a decreased superficial charge density weaken the hydrophobic interac- of SDS and alkyl polyglycosides in 4-hour epidermal patch tests. The au-
tion and electrostatic attraction of surfactants to the protein, which in thors performed a visual assessment of the solution application sites
turn makes the compounds milder to the skin. and correlated the results with the CMC values determined for the mix-
Generally, the literature describes a wide range of anionic surfac- tures. Even a small amount of a nonionic surfactant added to an SDS so-
tants which are considered as mild-acting [1,9,14,15]. Such compounds lution was found to decrease the irritant activity, compared to a 20%
include sulphosuccinates [180], isethionates [35,36], sarcosinates [155], solution of sodium lauryl sulphate. By comparing the results with the
acylated protein hydrolysates and their salts [181], alkyl ether carboxyl- CMC of the mixtures, the authors demonstrated a correlation between
ates [157], glutamic acid derivatives [38] and certain salts of alkyl ether the two parameters. A decrease in the micelle-forming concentration
sulphates (magnesium and zinc) [97,159]. ensures a reduction in the irritant activity of the surfactant solution
[182].
4.2. Application of surfactant mixtures Similar reports have been published by Kawasaki et al. [183] and Lee
et al. [184] presenting results of visual assessments and TEWL measure-
Surfactant mixtures are widely used in body wash cosmetics and ments after patch tests performed for mixtures of SDS and sodium
household chemicals on account of the synergistic effect achieved by lauroyl glutamate (SLG). A fall in the concentration of SDS in the solu-
different surfactants added to such mixtures [11,57–64]. For example, tion, coupled with a simultaneous increase in the SLG level, caused a de-
nonionic surfactants used in combination with an anionic surfactant in crease in TEWL and a drop in the visually assessed severity of the skin
the solution improves the detergent properties of the system. The appli- irritation effect. According to the authors, a decrease in the CMC of the
cation of an amphoteric compound enhances the stability of foam gen- SDS solution achieved by adding another type of anionic surfactant re-
erated in the solution of anionic surfactants [6,57,58,63–65]. The duces the concentration of free monomers in the volume phase, which
widespread use of mixtures of different surfactants has brought about limits interactions between the compounds and the proteins in the stra-
extensive studies investigating their skin irritation potential. As the lit- tum corneum. Another important aspect is the so-called effect of com-
erature demonstrates, the application of mixtures containing surfac- petition for contact sites which occurs between both compounds.
tants of diverse types is one of the simplest methods to increase the Tadenuma et al. [185] assessed the impact of mixtures containing
size of micelles in anionic surfactants and make them more stable. SDS and ethoxylated lauryl alcohol on bovine serum albumin. Based
Also, mixtures of different surfactants have an effect on lowering the on their results, they argue that the addition of a nonionic surfactant
CMC, which decreases the concentration of monomers in the volume to an aqueous SDS solution markedly reduced the denaturation of the
phase. Research shows that the above effect triggers a reduction in the protein – bovine serum albumin. The observed effect is stronger in pro-
skin irritation potential [1–5,13–15]. Fig. 4 shows a schematic represen- portion to the concentration of ethoxylate in the system. The authors
tation of changes to the structure of micelles formed in the solution con- explain the phenomenon by pointing out that the molecules of nonionic
taining mixtures of various surfactant types. surfactants are the first to become adsorbed on the protein chains,
As shown in the figure, micelles formed in the solution by the same which reduces the number of active centres available for SDS monomer
surfactant (e.g. anionic) are relatively small in size (A), which can pro- binding. Paye et al. [186] show that the denaturation of proteins is al-
mote their penetration via the skin barrier and impact on the structural ways more prominent in systems containing one type of surfactants
elements of the epidermis. Strong electrostatic repulsion of hydrophilic than in surfactant mixtures. The same conclusions have been presented
fragments in the micelle-forming molecules has an effect on reducing by Miyazawa et al. [187], showing that the application of surfactant
A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255 251

Fig. 4. Structure of micelles in surfactant mixtures: A – mixture of the same anionic surfactant, B – mixture of anionic and cationic surfactants, C – mixture of anionic and nonionic
surfactants, D – mixture of two different anionic surfactants, E – mixture of anionic and amphoteric surfactants. Own study based on [14].

mixtures has an effect on reducing the number of monomers in the so- studies warrant the conclusion that the application of mixtures of oppo-
lution. The author claims that monomers play main role in the develop- sitely charged surfactants may contribute to improving the mildness of
ment of skin irritations. Blake [33] and Paye [164] in their studies formulations (in the context of protein interactions) but, at the same
demonstrate that the incorporation of amphoteric surfactants into an time, it may increase solubilization processes affecting epidermal lipids
SDS solution markedly reduces the skin irritation potential by limiting which, as a consequence, may manifest as excessive skin dryness.
the possibilities for surfactant molecule binding to protein molecules. A number of researchers have described interactions between
Dominguez et al. [188] have studied SDS solutions with cocamidopropyl surfactants in their mixtures resulting in a decrease of the skin irritation
betaine, concluding that a lower skin irritation potential is associated potential in relation to the more irritant compound. Nearly all
with the use of two different types of surfactants forming mixed mi- authors point to a direct correlation between a decrease in the CMC of
celles that are much larger in size and more stable than those present surfactant mixtures and a reduction in the skin irritation potential
in the SDS solution. Also, McFadden et al. [189] have conducted epider- [13–16,160–167].
mal patch tests with mixtures of SDS and the cationic surfactant
benzalkonium chloride, noting that the addition of a cationic surfactant 4.3. Addition of polymers and biopolymers
to an anionic solution has an effect on changing the structures of aggre-
gates forming in the solution, which reduces its irritant activity com- Unique interactions occurring between polymers and surfactants
pared to pure SDS solutions. may lead to a decrease in the skin irritation potential. Specific
Interesting findings have been reported by Ananthapadmanabhan polymer-surfactant complexes developing in mixtures result in the sys-
et al. [131] based on comparisons between the results of in vitro tests tem having different properties than in polymer-free surfactant solu-
of protein denaturation (zein test) and extraction of intercellular lipids tions [13–15,17,58,191]. Research in this scope shows that the effect is
in the SC (stearic acid washing test) by aqueous solutions of ethoxylated characteristic both for synthetic polymers (polyvinylpyrrolidone [192],
sodium lauryl sulphate and cocamidopropyl betaine. Even though, in polyethylene glycols [185], acrylic derivatives [193]) and biopolymers
line with the authors' expectations, the zein number was found to de- (proteins and their hydrolyzed derivatives and rubbers [194,195–197]).
crease along with increasing amphoteric surfactant concentrations in The mechanisms responsible for surfactant and polymer interactions
the solution, the second test demonstrated more intensive lipid wash- in an aqueous solution are complex processes which give such a system
ing. The authors thus conclude that adding amphoteric or nonionic sur- different properties than in polymer-free systems [198–202]. An expla-
factants to a solution of anionic surfactants does not guarantee the nation for this phenomenon can be found by analyzing the relationship
attainment of systems that would be perfectly safe for the skin. The re- between surface tension as a function of surfactant concentration in
searchers note that the evaluation of the skin irritant effect must be polymer-enriched and polymer-free systems. Within the identified re-
comprehensive in scope, including interactions between surfactants lationship, a few regions can be noted which reflect the nature of poly-
and proteins and lipids in the SC. mer and surfactant interactions, and where the curve generated for the
Recently, interesting study results were published by Chen et al. relationship has a different shape compared to the additive-free aque-
[178] who investigated mixtures of anionic SDS with cationic Oleyl Bis ous surfactant solution [191,202]. An example of the relationship is
(2-hydroxyethyl)methyl Ammonium Bromide (OHAB) in contact with shown in Fig. 5.
the zein protein. Mixtures of these compounds were found to have At a low concentration of the surface active agent, monomers are
lower zein solubilization ability. The authors conclude that mixing op- adsorbed at the interface between phases (e.g. water – air). Polymer
positely charged surfactants reduces the net charge of mixed aggre- chains bound to the surfactant molecules are also adsorbed at the inter-
gates, which weakens the electrostatic attraction between the facial surface. In the volume phase of the solution, the surfactant con-
aggregates and zein. In addition, the large size of the arising aggregates centration is low, and polymer chains occur in the form of the so-
may increase their steric repulsion in relation to the zein backbone. The called flocs. Such tangled chains represent an additional interface be-
authors point out that it is possible to obtain a mixture with a low CMC tween phases, where molecules of the surface active agent are able to
value and high surface activity by appropriately adjusting its quantita- be adsorbed (area I in the chart). As the surfactant concentration in
tive and qualitative composition. Moreover, since large-sized aggre- the solution rises, the monomers continue to settle at the interface be-
gates formed in this manner are characterized by diminished zein tween phases, while interactions with water result in gradual disentan-
solubilization ability, their irritant effect is also reduced. The research glement of polymer chains. Monomers acquire an additional interface
is highly promising from the practical perspective, in the context of for- where they are able to be adsorbed [58,191,198–202].
mulation of cosmetic detergent systems. In another study, the same au- Further increases in the surfactant concentration result in the forma-
thor and co-workers [190] investigated mixtures of SLS with OHAB to tion of micelles bound to polymer chains. The phenomenon takes place
determine their ability to solubilize phospholipids and penetrate the at a specific concentration referred to as the critical aggregation concen-
skin. Based on the study findings, they conclude that aggregates formed tration (CAC). The correlations presented above demonstrate that the
in anionic/cationic surfactant mixtures have high lipid solubilization addition of a polymer to a surfactant solution accelerates the process
ability and a reduced skin penetration potential. The results of the two of aggregation of surface active agents at a concentration lower than
252 A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255

Fig. 5. Schematic representation of interactions between polymers and surfactants in aqueous solutions. Own study based on [106,198–203].

the critical micelle concentration (CMC). Depending on the structure of the principle, ionic interactions present in such systems are stronger
the polymer and the surfactant, a variety of interaction types can be than hydrophobic interactions, and hydrophobic interactions are supe-
observed between them. As Nagarajan shows [203], electrostatic, rior in strength to ion-dipole interactions. Factors influencing these pro-
hydrophobic and dipole-dipole interactions are observed, as well as cesses are linked both to the surface active agent and the polymer. The
hydrogen bonds between the macromolecule and the surfactants. The former include the structure of the hydrophobic and hydrophilic parts
formation of polymer-surfactant aggregates reduces the mobility of of the surfactant and the concentration of the surface active agent,
hydrophilic groups in the micelle, which translates into an increase in while the latter comprise the chemical nature, charge density, hydro-
aggregate stability. In addition, polymer chains have an ability to pene- phobicity as well as conformation and flexibility of the polymer chain.
trate partially into the structure of the micelle, increasing its size. An important role can also be played by the solvent on account of its ca-
As the surfactant concentration continues to rise, the desorption of pacity to affect the degree of dissociation of ionizable groups [212–227].
polymer chains from the surface phase to the volume phase is observed Between the CAC and CMC* points, polymers and surfactants may
(area II). In addition, a growing number of micelles become attached to form a new phase called “coacervate”, in which aggregates may grow
polymer chains. Moreover, an increase in the surfactant concentration in the process of self-association. However, it needs to be noted that a
and saturation of the interface cause the number of monomers in the strong electrostatic interaction in systems containing an oppositely
volume phase to grow. The saturation of the polymer – solution interfa- charged surfactant and polyelectrolyte may induce cooperative binding
cial surface by the micelles (resulting in the so-called “pearl-necklace” of surfactant molecules to the polyelectrolyte even at concentrations
structures, with the micelles wrapped by polymer chains) leads to an in- below the CAC point [228]. Similar observations were made by Purcell,
crease in the number of monomers in the volume phase of the solution, Lu and Thomas [229], who studied SDS solutions with polyvinylpyrrol-
where free polymer-unbound micelles ultimately develop (area III) at a idone. They found that below the critical aggregation concentration
characteristic surfactant solution. Their development, like in surfactant (CAC) the adsorption of SDS is increased by the presence of PVP, thus
solutions, is attributable to the hydrophobic effect [198,202,203]. The demonstrating that both components interact on the surface. Further-
concentration at which micelles begin to form in the volume phase in more, the authors observed the formation of polymer-surfactant com-
the polymer-surfactant system is often referred to as the apparent crit- plexes even below the CAC. Lange [230] concluded that the surface
ical micelle concentration (CMC*) [204]. tension of polymer-surfactant systems might be lower than the surface
Interactions between surfactants and polymers have been investi- tension of pure surfactant solutions, showing that the polymer or
gated in multiple scientific studies for many years. Research in this polymer-surfactant complex has surface active properties. It should
field plays an important role from the viewpoint of practical applica- also be noted that in systems containing a polyelectrolyte and a surface
tions in food, pharmaceutical and cosmetic industries [205–210]. active agent with opposite charges the relationship between surface
Depending on the molecular structures of the polymer and the tension and surfactant concentration may demonstrate more than
surface-active agent, and the nature of forces occurring between them, three transition points referred to above [231].
various types of polymer-surfactant interactions can be distinguished There are a number of methods which can be used to investigate in-
[203–210]. The main forces responsible for interactions taking place be- teractions taking place between polymers and surfactants in solutions.
tween the polymer and the surface active agent are electrostatic, dipole- The most important of them include surface tension, viscosimetry,
dipole and hydrophobic interactions, and hydrogen bonds between the calorimetry, nuclear magnetic resonance, conductometry, UV-VIS,
macromolecules of the polymer substance and surfactants. Bao et al. fluorescence and phosphorescence, potentiometry, vapour pressure
[211] formulated a general principle pertaining to forces occurring be- measurement, speed sound measurement, cyclic voltammetry and se-
tween polymers and surfactants in electrolyte solutions. According to lective electrodes for a given surfactant [232–243].
A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255 253

As shown, the incorporation of a polymer into a surfactant solution contain a hydrophobic substance, and an adsorptive surfactant layer
leads to the “uptake” of surfactant monomers in the volume phase and forming on their surface. An important factor is that the incorporation
furthermore increases the size and stability of micelles [191,198–204]. of a hydrophobic phase into a surfactant solution results in the forma-
The risk of skin irritation in such systems is minimized, as evidenced tion of an “additional” interface (between the hydrophobic and hydro-
by multiple studies [13,14,191–197]. philic phases), where monomers may be constituted. In this way, their
concentration in the system drops, resulting in the observed decrease
4.4. Addition of refatting and hydrophobic substances in the skin irritation potential. As the authors show, the extract used
in hand-dishwashing liquids can produce a positive skin effect not
Refatting substances, which are also defined as components rebuild- only on account of its hydrophobicity but also the content of com-
ing the skin barrier, are capable of reducing surfactant interactions with pounds with a proven medicinal benefit.
the hydrolipidic coat and intercellular lipids of the SC. Their activity
comprises the formation of an occlusive film on the skin, thus limiting 5. Conclusions
monomer access to the skin surface [244,245]. This activity, however,
requires repeated application of a product containing substances of As the discussion above reveals, the mechanism which underlies the
this type. In addition, the decrease in the skin irritation potential in- skin irritation potential of surfactants is complex, with an important role
duced by a surfactant solution may result from interactions between being played by the form in which the compound occurs in the aqueous
the compounds and micelles. By embedding into the aggregate struc- solution. The severity of the irritant effect is conditional on a number of
ture, they increase its size and improve stability. Substances of this factors such as the type, concentration and duration of skin exposure to
type include esters of fatty acids and their ethoxylated derivatives surfactants. Based on the literature review shown above, studies evalu-
(e.g. PEG-7 Glyceryl Cocoate, Sucrose Cocoate), ethoxylated glycerides ating methods of reducing the skin irritation potential of household
and triglycerides (e.g. PEG-60 Maracuia Glycerides) as well as lanolin chemicals and body wash cosmetics should mainly take into account
derivatives (mainly PEG-75 Lanolin). Based on the TEWL and skin hy- ways to reduce the concentration of monomers in the solution and in-
dration measurements, refatting additives have been shown to contrib- crease the size and stability of emerging micellar aggregates. The choice
ute to a substantial decrease in the loss of water from the epidermis, of an appropriate method of reducing the skin irritation potential, for
which is associated with a lower degree of damage to the skin barrier. example by using mixtures of different surfactants, adding polymers
They exhibit an ability to bind water in the skin themselves [13,14,65]. or refatting substances, should be based on a multifactorial analysis of
In recent years much attention has been paid to the application of the irritant effect, including surfactant interactions with epidermal pro-
hydrophobic compounds in body wash products, where they serve as teins and the cement and lipid film in the stratum corneum.
additives preventing the development of skin irritation. Research in
this area focuses on the application of plant oils, hydrophobic plant ex- Declarations of interest
tracts and fatty acids [245–249]. The most recent scientific develop-
ments suggest that the use of raw materials that are similar in None.
quantitative and qualitative composition to the ingredients naturally
occurring in the intercellular cement of the stratum corneum may com-
Acknowledgements
pensate losses of the components occurring in the washing process
[250–255]. Mukherjee et al. [255], studying polar sunflower oil and Funded with the assistance of the Ministry of Science and Higher
completely non-polar mineral oil, have found that the addition of oil Education from subsidies for statutory activity. Project no. 3086/35/P
markedly reduces skin irritation. The authors attribute the observed ef- entitled “Development of formulations and technologies for the manu-
fect to the fact that different oils variously interact with the proteins facture of innovative cosmetics, pharmacy supplies, household and in-
found in the stratum corneum. The same author has conducted a dustrial chemicals”.
study to determine the effect of stearic acid added to mild body wash
gels on the composition of the intercellular cement in the SC, comparing References
the findings with results obtained for products containing soybean oil
and vaseline. The study has shown that molecules of stearic acid from [1] Paye M. In: Chew AL, Maibach HI, editors. Irritant dermatitis. Berlin Heidelberg:
Springer; 2006 [Chapter 45].
the product can become embedded into the structure of the intercellu-
[2] Berardesca E, Distante F. Contact Dermatitis 1994;31:281–7.
lar cement within the SC, correcting its deficiency in the SC composition [3] Wolf R, Wolf D, Tüzün B, Tüzün Y. Clin Dermatol 2001;19:502–15.
due to the washing process [254]. Similar results have been reported by [4] Mehta SS, Reddy BS. Int J Dermatol 2003;42:533–42.
Conti et al. [253] following an analysis of the effects of triglycerides of [5] Abbas S, Goldberg JW, Massaro M. Dermatol Ther 2004;17(s1):35–42.
[6] Karsa DR. Industrial applications of surfactants IV. Elsevier; 1999.
fatty acids rich in linoleic acid radicals added to body wash products. [7] Rosen MJ, Zhu ZH, Gao T. J Colloid Interface Sci 1993;157(1):254–9.
Based on their studies, the authors conclude that the application of [8] Shiloach A, Blankschtein D. Langmuir 1998;14(7):1618–36.
products containing esters of linoleic acid can correct changes in the [9] Mehling A, Kleber M, Hensen H. Food Chem Toxicol 2007;45(5):747–58.
[10] Adamson AW. Physical chemistry of surfaces. 5th ed. New York: Wiley; 1990.
composition of the intercellular cement in the epidermis and signifi- [11] Piorr R. In: Falbe J, editor. Surfactants in consumer products: theory, technology
cantly improve the epidermal protective layer. and application. Springer Science & Business Media; 2012 [Chapter 1].
Wasilewski et al. [256,257] have analyzed how the concentration of [12] Sułek MW, Ogorzałek M, Wasilewski T, Klimaszewska E. J Surfactant Deterg 2013;
16(3):369–75.
a hydrophobic extract obtained under supercritical CO2 conditions af- [13] Polefka TG. In: Broze G, editor. Handbook of detergents. Part A: properties. Marcel
fects the safety-in-use of hand dishwashing liquids, noting a decrease Dekker Inc.; 1999 [Chapter 11].
in the irritant effect induced by these products. They also noted that [14] Jackson CT, Paye M, Maibach HI. In: Barel A, Paye M, Maibach HI, editors. Handbook
of cosmetic science and technology fourth edition. CRC Press; 2014 [Chapter 32].
the additives reduced transepidermal water loss and epidermal dryness. [15] Lips A, Ananthapadmanabhan KP, Vethamuthu M, Hua XY, Yang L, Vincent C, et al.
Based on the findings of their study, the authors propose a mechanism, In: Rhein L, editor. Surfactants in personal care products and decorative cosmetics
according to which hydrophobic plant extracts form aggregates in the third edition. CRC Press Taylor & Francis Group; 2007 [Chapter 9].
[16] Nielsen GD, Nielsen JB, Andersen KE, Grandjean P. Int J Occup Environ Health 2000;
volume phase of the washing bath. The surface of the aggregates pro-
6(2):138–42.
vides the adsorption area for surfactant monomers responsible for the [17] Otzen D. Biochim Biophys Acta 2011;1814(5):562–91.
irritant effect. The researchers claim that the presence of the hydropho- [18] Grammer NY, Fitzpatric JE, Jackson RL, Horton H, Damiano MA. J Am Acad Dermatol
bic phase may lead to a decrease in the number of surfactant monomers. 1996;35:258–60.
[19] Lee CH, Howard IM. In: Chew A, Maibach HI, editors. Irritant dermatitis. Berlin:
In the case of low hydrophobic phase concentrations, intramicellar sol- Springer-Verlag; 2006 [Chapter 30].
ubilization occurs, with micelles transforming into aggregates which [20] Goffin V, Paye M, Pierard GE. Contact Dermatitis 1995;33:38–41.
254 A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255

[21] Barel A, Paye M, Maibach HI. Handbook of cosmetic science and technology. Third [83] Walters RM, Mao G, Gunn ET, Hornby S. Dermatol Res Pract 2012;2012:1–9.
edition. CRC Press; 2009. [84] Bromberg L, Liu X, Wang I, Smith S, Schwicker K, Eller Z, et al. Colloids Surf B
[22] Simion FA. In: Chew A, Maibach HI, editors. Irritant dermatitis. Berlin: Springer- Biointerfaces 2017;157:366–72.
Verlag; 2006 [Chapter 53]. [85] Pinazo A. Colloids Surf B Biointerfaces 1996;8(1–2):63–72.
[23] Morrison Jr BM, Paye M. J Soc Cosmet Chem 1995;46:291–9. [86] Moore PN, Puvvada S, Blankschtein D. Langmuir 2003;19(4):1009–16.
[24] Zeidler U. J Soc Cosmet Chem Jpn 1986;20(1):17–26. [87] Deo N, Jockusch S, Turro NJ, Somasundaran P. Langmuir 2003;19(12):5083–8.
[25] Robinson MK, McFadden JP, Basketter DA. Contact Dermatitis 2001;45:1–12. [88] Ruso JM, Deo N, Somasundaran. Langmuir 2004;20(21):8988–91.
[26] Prottey C, Ferguson T. J Soc Cosmet Chem 1975;26(1):29–46. [89] Mehta SK, Bhasin KK, Kumar A. Colloids Surf A Physicochem Eng Asp 2009;346(1):
[27] Frosch PJ, Kligman AM. J Am Acad Dermatol 1979;1(1):35–41. 195–201.
[28] Hubbard AW, Moore LJ, Clothier RH, Sulley H, Rollin KA. Toxicol In Vitro 1994;8(4): [90] Mehta SK, Bhasin KK, Kumar A. J Colloid Interface Sci 2008;323(2):426–34.
689–91. [91] Kelley D, McClements DJ. Food Hydrocoll 2003;17(1):73–85.
[29] Pezron I, Galet L, Clausse DJ. Colloid Interface Sci 1996;180(1):285–9. [92] Cohen L, Sanchez E, Martin M, Soto F, Trujillo F. J Surfactant Deterg 2016;19(5):
[30] Götte E. Skin compatibility of tensides measured by their capacity for dissolving 1089–92.
zein protein. IVth international congress on surface active substances, Brussels, [93] Faucher JA, Goddard ED. J Soc Cosmet Chem 1978;29:323–37.
Belgium; 1964. p. 83. [94] Putterman GJ, Wolejsza NF, Wolfram MA, Laden K. J Soc Cosmet Chem 1977;28:
[31] Imokawa G, Sumura K, Katsumi M. J Am Oil Chem Soc 1975;52:484–9. 521–32.
[32] Tavss EA, Eigen E, Kligman AM. J Soc Cosmet Chem 1988;39:267–72. [95] Di Nardo A, Conti A, Martini M, Seidenari S. Skin Res Technol 1998;4:192–5.
[33] Blake-Haskins JC, Scala D, Rhein LD, Robbins CR. J Soc Cosmet Chem 1986;37: [96] Robbins CR, Fernee K. J Soc Cosmet Chem 1983;34:21–34.
199–210. [97] Cohen L, Martin M, Soto F, Trujillo F, Sanchez E. J Surfactant Deterg 2016;19(1):
[34] Grubauer G, Elias PM, Feingold KR. J Lipid Res 1989;30(3):323–33. 219–22.
[35] Ghosh S, Blankschtein D. Int J Cosmet Sci 2007;58(3):229–44. [98] Cohen L, Moreno A, Berna JL. J Surface Sci Technol 1997;13:59–71.
[36] Ananthapadmanabhan KP, Moore DJ, Subramanyan K, Misra M, Meyer F. Dermatol [99] Cohen L, Moreno A, Berna JL. Tenside Surfactant Deterg 1998;35:265–9.
Ther 2004;17(1):16–25. [100] Seweryn A, Wasilewski T, Bujak T. Ind Eng Chem Res 2016;55(4):1134–41.
[37] Takagi Y, Shimizu M, Morokuma Y, Miyaki M, Kiba A, Matsuo K, et al. Int J Cosmet [101] Menger FM, Littau CA. J Am Chem Soc 1993;115(22):10083–90.
Sci 2014;36(4):305–11. [102] Menger FM, Littau CA. J Am Chem Soc 1991;113(4):1451–2.
[38] Corazza M, Lauriola MM, Bianchi A, Zappaterra M, Virgili A. Dermatitis 2010;21(5): [103] Song LD, Rosen MJ. Langmuir 1996;12(5):1149–53.
262–8. [104] Acharya DP, Kunieda H, Shiba Y, Aratani KI. J Phys Chem B 2004;108(5):1790–7.
[39] Wickett RR, Visscher MO. Am J Infect Control 2006;34(10):98-10. [105] Shukla D, Tyagi VK. J Oleo Sci 2006;55(5):215–26.
[40] Ananthapadmanabhan KP, Mukherjee S, Chandar P. Int J Cosmet Sci 2013;35(4): [106] Rosen MJ, Tracy DJ. J Surfactant Deterg 1998;1(4):547–54.
337–45. [107] Zana R. Adv Colloid Interface Sci 2002;97(1):205–53.
[41] Rawlings AV, Scott IR, Harding CR, Bowser PA. J Invest Dermatol 1994;103(5): [108] Kamal MS. J Surfactant Deterg 2016;19(2):223–36.
731–40. [109] Kumar N, Tyagi R. Cosmetics 2013;1(1):3–13.
[42] Wertz PW, van der Bergh B. Chem Phys Lipids 1998;91:85–96. [110] Kumar N, Tyagi R. J Dispers Sci Technol 2014;35(2):205–14.
[43] Elias PM. J Invest Dermatol 1983;80(suppl):44–9. [111] Tsubone K, Ogawa T, Mimura K. J Surfactant Deterg 2003;6(1):39.
[44] Rawlings AV, Harding CR. Dermatol Ther 2004;17:43–8. [112] Diz M, Manresa A, Pinazo A, Erra P, Infante M. J Chem Soc Perkin Trans 1994;2(8):
[45] Verdier-Sévrain S, Bonté F. J Cosmet Dermatol 2007;6:75–82. 1871–6.
[46] Braun-Falco O, Korting HC. Hautarzt 1986;37(3):126–9. [113] Mulligan CN. Environ Pollut 2005;133(2):183–98.
[47] Baranda L, González-Amaro R, Torres-Alvarez B, Alvarez CM, Ramírez V. Int J [114] Banat IM. Bioresour Technol 1995;51(1):1–12.
Dermatol 2002;41(8):494–9. [115] Allemann IB, Baumann L. Skin Therapy Lett 2008;13(7):5–9.
[48] Korting HCh, Braun-Falco O. Clin Dermatol 1996;14:23–7. [116] Bai L, McClements DJ. J Colloid Interface Sci 2016;479:71–9.
[49] Lémery E, Briançon S, Chevalier Y, Bordes C, Oddos T, Gohier A, et al. Colloids Surf A [117] Vecino X, Cruz JM, Moldes AB, Rodrigues LR. Crit Rev Biotechnol 2017;37(7):
Physicochem Eng Asp 2015;469:166–79. 911–23.
[50] Danov KD, Kralchevsky PA, Ananthapadmanabhan KP. Adv Colloid Interface Sci [118] Ferreira A, Vecino X, Ferreira D, Cruz JM, Moldes AB, Rodrigues LR. Colloids Surf B
2014;206:17–45. Biointerfaces 2017;155:522–9.
[51] Kha A, Marques EF. Curr Opin Colloid Interface Sci 1999;4(6):402–10. [119] Lee S, Lee J, Yu H, Lim J. Colloids Surf A Physicochem Eng Asp 2018;536:224–33.
[52] Garcia MT, Ribosa I, Sanchez Leal J, Comelles F. J Am Oil Chem Soc 1992;69:25–9. [120] Burgos-Díaz C, Martín-Venegas R, Martínez V, Storniolo CE, et al. Int J Pharm 2013;
[53] Sein A, Engberta J. Langmuir 1995;11:455–65. 453(2):433–40.
[54] Sein A, Engberta J, Linden E, Pas JC. Langmuir 1993;9:1714–20. [121] Kim HS, Jeon JW, Kim SB, Oh HM, Kwon TJ, Yoon BD. Biotechnol Lett 2002;24(19):
[55] Lasic D. Biochem J 1988;256(1):1–11. 1637–41.
[56] Hunter RJ. Zeta potential in colloid science: principles and applications. Academic [122] Alonso C, Lucas R, Barba C, Marti M, Rubio L, et al. J Pharm Pharmacol 2015;67(7):
Press; 2013. 900–8.
[57] Karsa DR, Farn RJ. Chemistry and technology of surfactants. John Wiley & Sons; [123] Takahashi M, Morita T, Fukuoka T, Imura T, Kitamoto D. J Oleo Sci 2012;61(8):
2008 [Chapter 1]. 457–64.
[58] Rosen MJ, Kunjappu JT. Surfactants and interfacial phenomena. 4th ed. John Wiley [124] Luna JM, Rufino RD, Sarubbo LA, Rodrigues LR, Teixeira JA, de Campos-Takaki GM.
& Sons; 2012. Curr Microbiol 2011;62(5):1527–34.
[59] Maibaum L, Dinner AR, Chandler D. J Phys Chem B 2004;108(21):6778–81. [125] Gudina EJ, Teixeira JA, Rodrigues LR. Colloids Surf B Biointerfaces 2010;76(1):
[60] Kronberg B. Curr Opin Colloid Interface Sci 2016;22:14–22. 298–304.
[61] Penfold J, Tucker I, Thomas RK, Staples E, Schuermann R. J Phys Chem B 2005;109 [126] Madhu AN, Prapulla SG. Appl Biochem Biotechnol 2014;172(4):1777–89.
(21):10760–70. [127] Gudiña EJ, Fernandes EC, Teixeira JA, Rodrigues LR. RSC Adv 2015;5(110):90960–8.
[62] Goloub TP, Pugh RJ, Zhmud BV. J Colloid Interface Sci 2000;229(1):72–81. [128] Imokawa G, Akasaki S, Minematsu Y, Kawai M. Arch Dermatol Res 1989;281:
[63] Gambogi J, Arvanitidou ES, Lai KY. In: Lai KY, editor. Liquid detergents. Second edi- 45–51.
tion. Taylor & Francis Group; 2006 [Chapter 1]. [129] Froebe CL, Simion FA, Rhein LD, Cagan RH, Kligman A. Dermatology 1990;181:
[64] Showell MS. Handbook of detergents. Part D: formulations. Taylor & Francis Group; 277–83.
2006. [130] Fulmer AW, Kramer GJ. J Invest Dermatol 1986;5:598–602.
[65] Gallegos C, Franco JM. Curr Opin Colloid Interface Sci 1999;4(4):288–93. [131] Ananthapadmanabhan KP, Yang L, Vincent C, et al. Quadrant 2009;22(6):
[66] Rhein LD, Simion FA. Ser Surf Sci 1991;32:33–49. 307–16.
[67] Rhein L, Robbins C, Fernee K, Cantore R. J Soc Cosmet Chem 1986;37:125–39. [132] Lémery E, Briançon S, Chevalier Y, Bordes C, Oddos T, Gohier A, et al. Colloids Surf A
[68] Wilhelm KP, Cua AB, Wolff HH, Maibach HI. J Invest Dermatol 1994;101:310–5. Physicochem Eng Asp 2015;469:166–79.
[69] Wihelm KP. Curr Probl Dermatol 1995;22:72–9. [133] Som I, Bhatia K, Yasir M. Status of surfactants as penetration enhancers in transder-
[70] Ciurleo A, Cinelli S, Guidi M, Bonincontro A, Onori G, LaMesa C. Biomacromolecules mal drug delivery. J Pharm Bioallied Sci 2012;4(1):2–9.
2007;8(2):399–405. [134] Astner S, Gonzalez E, Cheung AC, Rius-Diaz F, Doukas AG, et al. J Invest Dermatol
[71] Stenstam A, Topgaard D, Wennerström H. J Phys Chem B 2003;107(32):7987–92. 2005;124:351–9.
[72] Goddard ED, Ananthapadmanabhan KP. Interactions of surfactants with polymers [135] Lee JK, Kim DB, Kim JI, Kim PY. Toxicol In Vitro 2000;14(4):345–9.
and proteins. CRC Press; 1993. [136] Effend I, Maibach HI. Contact Dermatitis 1995;33(4):217–25.
[73] Moriyama Y, Takeda K. Langmuir 2005;21(12):5524–8. [137] Corazza M, Lauriola MM, Zappaterra M, Bianchi A, Virgili A. J Eur Acad Dermatol
[74] Giroux HJ, Britten M. Food Hydrocoll 2004;18(4):685–92. Venereol 2010;24(1):1–6.
[75] Ghosh S. Colloids Surf A Physicochem Eng Asp 2005;264(1):6–16. [138] Nosbaum A, Vocanson M, Rozieres A, Hennino A, Nicolas JF. Eur J Dermatol 2009;19
[76] Ghosh S. Colloids Surf B Biointerfaces 2008;66(2):178–86. (4):325–32.
[77] De Jongh CM, Verberk MM, Withagen CET, Jacobs JJL, Rustemeyer T, Kezic S. Con- [139] Slodownik D, Lee A, Nixon R. Aust J Dermatol 2008;49:1–11.
tact Dermatitis 2006;54:325–33. [140] Frosch PJ, John SM. In: Johansen JD, Frosch PJ, Lepoittevin JP, editors. Contact
[78] Mackie A, Wilde P. Adv Colloid Interface Sci 2005;117(1):3–13. dermatitis. Berlin Heidelberg: Springer; 2011 [Chapter 15].
[79] Jones MN. Chem Soc Rev 1992;21(2):127–36. [141] Berardesca E, Distante F. Contact Dermatitis 1994;31(5):281–7.
[80] Fluhr JW, Akengin A, Bornkessel A, Fuchs S, Praessler J, et al. Br J Dermatol 2005; [142] Yokota M, Maibach HI. Contact Dermatitis 2006;55(2):65–72.
143:125–31. [143] Levin CY, Maibach HI. Int J Immunopharmacol 2002;2:183–9.
[81] Moore PN, Puvvada S, Blankschtein D. J Cosmet Sci 2003;54(1):29–46. [144] Van Der Valk PG, Nater JP, Bleumink E. J Invest Dermatol 1984;82(3):291–3.
[82] Walters RM, Fevola MJ, Librizzi JJ, Martin K. Cosmet Toilet 2008;123(12):53–60. [145] Zhai H, Maibach HI. Contact Dermatitis 2001;44(4):201.
A. Seweryn / Advances in Colloid and Interface Science 256 (2018) 242–255 255

[146] Belsito DV, Fransway AF, Fowler JF, Sherertz EF, Maibach HI, Mark JG, et al. J Am [203] Nagarajan R. Polymer–surfactant interactions. New horizons: detergents for the
Acad Dermatol 2002;46(2):200–6. new millennium conference invited papers. FortMyers, Florida: American Oil
[147] Groot AC, Walle HB, Weyland JW. Contact Dermatitis 1995;33(6):419–22. Chemists Society and Consumer Specialty Products Association; 2001.
[148] Blondeel A. Contact Dermatitis 2003;49(6):304–5. [204] Jones MN. J Colloid Interface Sci 1967;23(1):36–42.
[149] Gijbels D, Timmermans A, Serrano P, Verreycken E, Goossens A. Contact Dermatitis [205] Goddard ED, Phillips TS, Hannan RB. J Soc Cosmet Chem 1975;26:461–75.
2014;70(3):175–82. [206] Goddard ED, Ananthapadmanabhan KP, editors. Interactions of surfactants with
[150] Andrade P, Gonçalo M, Figueiredo A. Contact Dermatitis 2010;62(2):119–20. polymers and proteins. Boca Raton, FL: CRC Press; 1993.
[151] Malanin KEN. Contact Dermatitis 2002;47(1):50. [207] Goddard ED, Ananthapadmanabhan KP, editors. Applications of polymer–
[152] Biebl KA, Warshaw EM. Dermatol Clin 2006;24(2):215–32. surfactant systems. Surfactant science seriesMarcel Dekker; 1998.
[153] Wolf R, Wolf D, Tüzün B, Tüzün Y. Dermatol Ther 2001;14:181–7. [208] Taylor DJF, Thomas RK, Penfold J. Adv Colloid Interface Sci 2007;132(2):69–110.
[154] Thomson KF, Wilkinson SM. Br J Dermatol 2000;142:84–8. [209] Nylander T, Samoshina Y, Lindman B. Adv Colloid Interface Sci 2006;123:105–23.
[155] Ray GB, Ghosh S, Moulik SP. J Surfactant Deterg 2009;12(2):131–43. [210] JCT Kwak, editor. Polymer-surfactant systems. Surfactant science seriesNew York:
[156] Takagi Y, Shimizu M, Morokuma Y, Miyaki M, et al. Int J Cosmet Sci 2014;36(4): Marcel Dekker; 1998.
305–11. [211] Bao H, Li L, Gan LH, Zhang H. Macromolecules 2008;41:9406–12.
[157] Ozawa T, Endo K, Masui T, Miyaki M, Matsuo K, Yamada S. J Surfactant Deterg 2016; [212] Vinceković M, Bujan M, Šmit I, Filipović-Vinceković N. Colloids Surf A Physicochem
19(4):785–94. Eng Asp 2005;255:181–91.
[158] Corazza M, Lauriola MM, Bianchi A, Zappaterra M, Virgili A. Dermatitis 2010;21(5): [213] Bakshi MS, Kaur I. Colloids Surf A Physicochem Eng Asp 2003;224:185–97.
262–8. [214] Thalberg K, Lindman B, Karlstrom G. J Phys Chem 1991;95:3370–6.
[159] Bigotti C, Guaio F, Merlo E, Gazzanig G, Villa G. J Appl Cosmet 2005;23:139–47. [215] Taylor DJF, Thomas RK, Hines JD, Humphreys K, Penfold J. Langmuir 2002;18:
[160] Blake-Haskins J, Scala D, Rhein L, Robbins C. J Soc Cosmet Chem 1985;36:379. 9783–91.
[161] Rhein L. In: Johansson I, Somasundaran P, editors. Handbook for cleaning/decon- [216] Hayakawa K, Kwak JCT. J Phys Chem 1983;87:506–9.
tamination of surfaces. Elsevier; 2007 [Chapter C3]. [217] Hayakawa K, Santerre JP, Kwak JCT. Macromolecules 1983;16:1642–5.
[162] Basketter DA, Marriott M, Gilmour NJ, White IR. Contact Dermatitis 2004;50(4): [218] Hayakawa K, Kwak JCT. J Phys Chem 1982;86:3866–70.
213–7. [219] Bai G, Wang Y, Yan H, Thomas RK, Kwak JCT. J Phys Chem B 2002;106:2153–9.
[163] Zhou C, Wang D, Cao M, Chen Y, Liu Z, Wu C, et al. ACS Appl Mater Interfaces 2016; [220] Thalberg K, Lindman B, Karlström G. J Phys Chem 1990;94:4289–95.
8(45):30811–23. [221] Thalberg K, Lindman B. J Phys Chem 1989;93:1478–83.
[164] Li Y, Wang X, Wang Y. J Phys Chem B 2006;110(16):8499-05. [222] Cardenas M, Nylander T, Joensson B, Lindman B. J Colloid Interface Sci 2005;286:
[165] Bakshi MS, Singh J, Kaur G. Chem Phys Lipids 2005;138(1–2):81–92. 166–75.
[166] Blagojević SN, Blagojević SM, Pejić ND. J Surfactant Deterg 2016;19(2):363–72. [223] Cooke DJ, Dong CC, Lu JR, Thomas RK, Simister EA, Penfold J. J Phys Chem B 1998;
[167] Blagojević SM, Pejić ND, Blagojević SN, Russ J. Phys Chem A 2017;91(13):2690–5. 102:4912–7.
[168] Jones MN. Biochem J 1975;151(1):109–14. [224] Alves L, Lindman B, Klotz B, Böttcher A, Haake HM, Antunes FE. J Colloid Interface
[169] Jones MN, Manley P. J Chem Soc Faraday Trans I 1979;75:1736–44. Sci 2018;513:489–96.
[170] Subramanian M, Sheshadri BS, Venkatappa MP. J Biosci 1986;10:359–71. [225] Thomas RK, Penfold J. Curr Opin Colloid Interface Sci 1996;1(1):23–33.
[171] Reynolds J, Herbert S, Steinhardt J. Biochemistry 1968;7:1357–61. [226] Penfold J, Thomas RK, Taylor DJF. Curr Opin Colloid Interface Sci 2006;11:337–44.
[172] Kaneshina S, Tanaka M, Kondo T, Mizuno T, Aoki K. Bull Chem Soc Jpn 1973;46: [227] Persson B, Hugerth A, Caram-Lelham N, Sundelof LO. Langmuir 2000;16:313–7.
2735–8. [228] Babak VG, Vikhoreva GA, Lukina IG. Colloids Surf A Physicochem Eng Asp 1997;
[173] Reynolds JA, Gallagher JP, Steinhardt J. Biochemistry 1970;9:1232–8. 128:75–89.
[174] Sarmiento F, Ruso JM, Prieto G, Mosquera V. Langmuir 1998;14:5725–9. [229] Purcell IP, Lu JR, Thomas RK, Howe AM, Penfold J. Langmuir 1998;14(7):1637–45.
[175] Takeda K, Sasa K, Kawamoto K, Wada A, Aoki K. J Colloid Interface Sci 1988;124: [230] Lange H. Kolloid Z Z Polym 1971;243:101–9.
284–9. [231] Bahramian A, Thomas RK, Penfold J. J Phys Chem B 2014;118:2769–83.
[176] Reynolds JA, Tanford C. Proc Natl Acad Sci 1970;66:1002–7. [232] Sehgal P, Mizuki T, Doe H, Wimmer R, Larsen KL, Otzen DE. Colloid Polym Sci 2009;
[177] James-Smith MA, Hellner B, Annunziato N, Mitragotri S. Ann Biomed Eng 2011;39: 287:1243–52.
1215–23. [233] Guo R, Zhu XJ, Guo X. Colloid Polym Sci 2003;281:876–81.
[178] Chen Y, Ji X, Han Y, Wang YL. Langmuir 2016;32:8212–21. [234] Okubo T, Kitano H, Ise N. J Phys Chem 1976;80:2661–4.
[179] Ananthapadmanabhan KP, Yu KK, Meyers CL, Aronson MP. J Soc Cosmet Chem [235] Zhou Y, Yu H, Zhang L, Sun J, Wu L, Lu Q, et al. J Incl Phenom Macrocycl Chem 2008;
1996;47:185–200. 61:259–64.
[180] Tyagi VK. J Oleo Sci 2006;55(9):429–39. [236] Valente AJM, Dinis CJS, Pereira RFP, Ribeiro ACF, Lobo VMM. Port Electrochem Acta
[181] Secchi G. Clin Dermatol 2008;26(4):321–5. 2006;24:129–36.
[182] Hall-Manning TJ, Holland GH, Rennie G, Revell P, Hines J, Barrat MD, et al. Food [237] Rafati AA, Safatian F. Phys Chem Liq 2008;46:587–98.
Chem Toxicol 1998;36:233–8. [238] Li J, Zhao C, Chao J. J Incl Phenom Macrocycl Chem 2008;62:325–31.
[183] Kawasaki Y, Quan D, Sakamoto K, Cooke R, Maibach HI. Skin Res Technol 1999;5 [239] Fonasaki N, Ishikawa S, Hirota S. Anal Chim Acta 2006;555:278–85.
(2):96–101. [240] Tang B, Wang X, Wang G, Yu Ch, Chen Zh. Talanta 2006;69:113–20.
[184] Lee CH, Kawasaki Y, Maibach HI. Contact Dermatitis 1994;30:205–9. [241] Mwakibete H, Cristantino R, Bloor DM, Wyn-Jones E, Holzwarth JF. Langmuir 1995;
[185] Tadenuma H, Yamada K, Tamura T. Jpn Oil Chem Soc 1999;48:207–13. 11:57–60.
[186] Paye M, Block C, Hamaide N, Hüttman GE, Kirkwood S, et al. Tenside Surfactant [242] Dharmawardana UR, Christian SD, Tucker EE, Taylor RW, Scamehorm JF. Langmuir
Deterg 2006;43(6):290–4. 1993;9:2258–63.
[187] Miyazawa K, Ogawa M, Mitsui T. Int J Cosmet Sci 1984;6:33–46. [243] Wilson LD, Verrall RE. J Phys Chem B 1998;102:480–8.
[188] Dominguez JG, Balaguer F, Parra JL, Pelejero CM. Int J Cosmet Sci 1981;3: [244] Elder R. L. J Am Acad Dermatol 1984;11(6):1168–74.
57–68. [245] Parente ME, Solana G. Int J Cosmet Sci 2005;27(6):354.
[189] McFadden JP, Holloway DB, Whittle EG, Basketter DA. Contact Dermatitis 2000;43 [246] Lodén M, Am J. Clin Dermatol 2003;4(11):771–88.
(5):264–6. [247] Levi K, Kwan A, Rhines AS, Gorcea M, Moore DJ, Dauskardt RH. Br J Dermatol 2010;
[190] Chen Y, Qiao F, Fan Y, Han Y, Wang Y. Langmuir 2017;33(11):2760–9. 163:695–703.
[191] Holmberg K, Jonsson B, Kronberg B, Lindman B. Surfactants and polymer in aque- [248] Lodén M, Buraczewska I, Edlund F. Br J Dermatol 2004;150(6):1142–7.
ous solution. John Wiley & Sons; 2002. [249] Yang B, Bonfigli A, Pagani V, Isohanni T, von Knorring A, Jutila A, et al. J Appl Cosmet
[192] Bujak T, Wasilewski T, Nizioł-Łukaszewska Z. Colloids Surf B Biointerfaces 2015; 2009;27(1):1–13.
135:497–503. [250] Förster T, Issberner U, Hensen H. J Surfactant Deterg 2000;3(3):345.
[193] Draelos Z, Hornby S, Walters RM, Appa Y. J Cosmet Dermatol 2013;12(4):314–21. [251] Takagi Y, Nakagawa H, Higuchi K, Imokawa G. Dermatology 2005;211:128–34.
[194] Goddard ED, Leung PS. Cosmet Toilet 1982;97:55–69. [252] Lee YB, Park HJ, Kwon MJ, Jeong SK, Cho SH. Eur J Dermatol 2011;21:710–6.
[195] Pugliese P, Hines G, Wielenga W. Cosmet Toilet 1990;105:105–11. [253] Conti A, Rogers J, Verdejo P, Harding CR, Rawlings AV. Int J Cosmet Sci 1996;18:
[196] Teglia A, Secchi G. Int J Cosmet Sci 1994;16:235–46. 1–12.
[197] Teglia A, Mazzola G, Secchi G. Cosmet Toilet 1993;108(11):56–65. [254] Mukherjee S, Edmunds M, Lei X, Ottaviani MF, Ananthapadmanabhan KP, Turro NJ.
[198] Goddard ED. Colloids Surf 1986;19(2–3):255–300. J Cosmet Dermatol 2010;9:202–10.
[199] Hansson P, Lindman B. Curr Opin Colloid Interface Sci 1996;1(5):604–13. [255] Mukherjee S, Yang L, Vincent C, Lei X, Ottaviani MF, Ananthapadmanabhan KP. Int J
[200] Loyen K, Iliopoulos I, Audebert R, Olsson U. Langmuir 1995;11(4):1053–6. Cosmet Sci 2015;37(4):371–8.
[201] Persson K, Bales BL. J Chem Soc Faraday Trans 1995;91(17):2863–70. [256] Wasilewski T, Seweryn A, Krajewski M. J Surfactant Deterg 2016;19(6):1315–26.
[202] Goddard ED, Hannan RB. J Am Oil Chem Soc 1977;54(12):561–6. [257] Wasilewski T, Seweryn A, Bujak T. Green Chem Lett Rev 2016;9(2):114–21.

You might also like