You are on page 1of 22

In: Detergents: Types, Components and Uses ISBN: 978-1-61668-986-5

Editors: Emilie T. Hagen © 2009 Nova Science Publishers, Inc.

Chapter 5

-SULFO FATTY METHYL ESTER


SULFONATES (-MES): A NEW
ANIONIC SURFACTANT

León Cohen1* David W. Roberts2


and Claudio Pratesi3
1
Escuela Politécnica Superior de Algeciras. Universidad de
Cádiz Avda Ramon Puyol s/n. 12202 Algeciras. Spain
2
Liverpool John Moores University
3
Sasol Italy S.p.A

ABSTRACT
-Sulfo Fatty Methyl Esters sulfonates (- MES) are new anionic
surfactants obtained via sulfoxidation of fatty acid methyl esters (FAME)
with SO2, O2, and ultraviolet light of appropriate wavelength. The
designation of Φ refers to the random positioning of SO3 in the alkyl
chain.
In this work we summarize the most relevant results of our research
started fifteen years ago, on the synthesis, separation, analysis and
performance of -sulfo fatty methyl ester sulfonates known as -MES,
and we update our last findings . We have to point out that we are for the

* Correspondance author: E-mail: leon.cohen@uca.es Ramón Puyol s/n 11202 - Algeciras (


Spain ).
2 León Cohen David W. Roberts and Claudio Pratesi

time being, the only research group in the world to have published our
research on -MES.
This work describes the optimum batch conditions for the
sulfoxidation of FAME and a reaction mechanism is proposed. The paper
depicts an improved workup for the separation of reaction products from
non reacted methyl ester and the GC-MS analysis of - sulfo fatty
methyl ester sulfonate is shown. Besides, an interpretation of conversion
and selectivity of sulfoxydation reaction is given.
Finally, performance of water solutions based on sulfoxylated methyl
ester of palmitic acid (-MES C16) have been studied and compared to
two leading types of surfactants used today: linear alkylbezene sulfonate
(LAS) secondary alkane sulfonate (SAS) and to -sulfo fatty methyl ester
sulfonate (-MES) with regard to solubility, performance and skin
compatibilty. The experimental results obtained indicate that -MES
can be regarded as a potential component of detergent formulations and
most likely of body care products.

INTRODUCTION
Fatty acid methyl esters are products derived from fats. The chain length
is within 12 to 18 carbon atoms. The names of the detergent range are
lauric(C12), myristic (C14), palmitic (C16) and stearic (C18).

O
CH3-CH2-…-CH2-…-CH2-C
 OCH3

α-Sulfo fatty methyl esters sulfonates (α-MES) are well known surfactants
derived from natural fats and oils, obtained via traditional sulfonation with
SO3. During recent years there have been a number of studies on their surface
properties, calcium tolerance, detergency and environmental behavior [1–5].
However, one of the distinct limitations in the use of α-MES is their poor
water solubility, which makes them less satisfactory products compared to
alkylaryl sulfonates (LAS).
α-MES exhibit a remarkable specificity in that sulfonate group is added
almost exclusively to the α position of the long chain when conventional
processes using SO3 is employed, presumably being the reason for their low
water solubility. To overcome this drawback some authors [6,7] suggested the
use of randomly sulfonated FAME, produced by ultraviolet (UV) irradiation–
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 3

sulfoxidation that apparently would eliminate the performance gap between


MES and LAS.
Sulfoxidation or sulfoxylation processes have already been used in the
detergent industry to produce paraffin secondary sulfonate (SAS) [8,9].
Sulfoxidation is the addition of sulphur dioxide and oxygen in the presence of
UV light of appropiate wavelength to some organic compounds, such as
aliphatic hydrocarbons or fatty acid methyl esters.
As a result of our investigation on -MES started in 1993, several papers
have been published [10-17]. These studies have led to two patents accepted in
1999 [18] and 2006 [19] by the Spanish Patent and Trade Mark Office.
The sulfoxydation reaction [14] proceeds through a radical mechanism
unlike the -MES traditional sulfonation, which proceeds through an
electrophilic substitution.
Due to the former, a characteristic of these compounds is that the
sulfonate group is randomly distributed over all , except the α and ω positions,
of the alkyl chain, being this the reason why they are named Φ- MES, as we
will see later.
The main reaction product is monosulfonic acid:

O
CH3-CH2-…-CH-…-CH2-C
OCH3
SO3H

Φ- MES Sulfonic Acid

The neutralization of the acid with sodium hydroxide gives the sodium
sulfonate or sulfoxylate, that belongs to the anionic surfactant family

CH3-CH2-…-CH-…-CH2- C
OCH3
SO3Na

Φ- MES Sodium Sulfonate

In this investigation the batch reaction was conducted in an anhydrous


medium, the optimal wavelength was determined and -MES were
4 León Cohen David W. Roberts and Claudio Pratesi

synthesized, separated and analyzed. Concerning performance, -MES


exhibited [11,17] properties that make them attractive as anionic surfactants,
namely:

 Good water solubility, that make them easy to be included in liquid


formulations.
 Very low viscosity of their aqueous solutions, that make them easy to
handle and pump.
 Very good wetting power
 Excellent water hardness stability, that alllow them to be formulated
in hard water regions.
 Excellent skin compatibility, that make them potentially very good
for hand dishwashing formulations and most likely for body care
products.
 Expected excellent biodegradation.
 Good detergent power.

In a previous paper [10], performances of -MES with varying chain


lengths were compared. Two main conclusions were drawn: the palmitic -
MES (-MES C16) could be used as LAS partner in detergent formulations
and the C16 alkyl chain was the optimal length. Since then, progress has been
made concerning the development of a procedure for separation of reaction
products [16] and analysis [12,13]. These investigations have led to products
with less impurities.
In the present work, we present the results of standard experimental tests,
that have allowed us to compare the performance of the most common
commercial anionic surfactants such as LAS (Linear Alkylbenzene Sulfonate),
SAS (Secondary Alkane Sulfonate) and -MES (-sulfo fatty Methyl Ester
Sulfonate) to -MES C16.

2. EXPERIMENTAL

Materials

Methyl esters were products from Procter & Gamble Chemicals, USA.
Gas chromatographic data are shown in Table 1.
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 5

Table 1.

Composition C12 C14 C16 C18


wt%
C12 98.9 1.2
C14 1.1 98.4 1.4
C16 0.4 98.6 2.4
C18 97.6

 LAS sodium salt. A commercial sodium sulfonate sample derived


from Petresul 550 acid manufactured by Petroquímica Española,
Spain.
 SAS sodium salt. A commercial cut of C14-C17 sample of Hostapur
SAS 93 from Clariant Germany.
 -MES C16-C18. A commercial sample of C16-C18 (50/50 wt%)
concentrated powder with 97% monosulfonate purity, from Desmet
Ballestra, Italy.

Photochemical Reactor and Sulfoxidation Procedure

All experiments were water free batch reactions. Experimental setup


consisted of a Rayonnet photochemical reactor with 16 lamps (Southern New
England Ultraviolet Company, Hamden, CT) . Three sets of lamps with three
different wavelengths (253.4, 300 and 350 nm) were assayed.
Operating conditions were as follows:

 Reaction temperature: 40ºC


 Volume of ME: 250 ml
 Gas flow: SO2 and O2 (excess)

Different reaction times were used ranging from 0,5 to 6 hours. At the end
of the reaction time, SO2 and O2 injections were stopped and N2 was then used
for 1 h to remove residual SO2.
6 León Cohen David W. Roberts and Claudio Pratesi

Separation Procedure

The main reaction components were monosulfonic acid, polysulfonic


acids and a very small amount of fatty acids coming from the hydrolisis of
methyl esters. Two procedures for the separation of the reaction products from
the non reacted methyl esters, are depicted in ref. 16 namely the water and the
hexane extraction.
As depicted in [16] in the hot water extraction, a volume of hot water
(250 mL at 60ªC ) was added to the reactor outlet in a separatory funnel. The
mixture was then shaken and both phases immediately separated . The upper
layer ( organic phase) contained all of the non reacted methyl ester and a very
small amount of fatty acid. The aqueous phase that is the lower layer,
contained all the sulfonic acid and a small amount of a mixture of fatty acid
and some residual methyl ester. The latter was then extracted with hexane in a
separatory funnel (100 mL is sufficient). Finally the aqueous phase was
neutralized with a sodium hydroxide solution. An updated network, has been
implemented.

Update. According to the water procedure described in ref.16, after the


separation of the non-reacted methyl ester, a small amount of dissolved methyl
ester and some fatty acid remained in the aqueous phase, that had to be
removed in order to improve the final product quality. An updated procedure,
is herein depicted. As mentioned in reference 16, the reactor outlet was mixed
with an equivalent amount of water and the two phases let to separate at 60ºC
in the oven. After separation of both phases, the unreacted methyl ester could
be used for a new run and the water solution followed further treatment
consisting in adding a volume of methanol equal to that of water. The mixture
was then extracted with a non polar solvent such as n-hexane. In the hexane
phase all the dissolved fatty acid and methyl ester were recovered, while the
water-methanol solution was distilled under a slightly reduced pressure and all
the methanol recovered. The remaining water solution was then neutralized
with sodium hydroxide and concentrated as desired. As mentioned above,
basically, the modification consisted of the use of a mixture of methanol-water
50/50 v/v and hexane, to perform the extraction of dissolved impurities.
Besides this, an improvement herein depicted, is that the upper layer
formed by the non reacted methyl ester can be reesterified with methanol in
order to convert the whole free fatty acid present. The improved detailed
procedure is depicted below in Scheme 1.
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 7

REACTOR OUTLET

Hot water Separation

Methanol Water Phase Non reacted


addition Methyl ester
Esterification

Hexane Extraction

Methanol-Water Hexane Phase


Phase

Hexane recovery
Methanol recovery
Neutralization

Scheme 1. Analytical Scheme work

3. RESULTS

Proposed Reaction Mechanism

We have summarized in Scheme 2 the ways in which the hydrocarbon


sulfoxidation reaction proceeds in the absence of water. As depicted in scheme
2, in the first reaction step the persulfonic acid molecule RSO4H is formed:

SO2 O2 RH
R. RSO2. RSO4. RSO4H + R.

Once formed, the persulfonic acid molecule RSO4H seems to proceed


through two mechanisms: a dark reaction, where the persulfonic molecule
disappears by thermal decomposition and an irradiated reaction, where the
persulfonic acid molecule RSO4H disappears through its action as an electron
donor.
8 León Cohen David W. Roberts and Claudio Pratesi

Thermal decomposition: Dark reaction


Thermal decomposition, corresponds to the pathway depicted in the lower
part of Scheme 2. Two new free radicals are generated by thermal
decomposition of the persulfonic acid molecule. However, not all the
persulfonic acid follows this pathway. Water formed by reaction of RH with
OH radical, converts part of the persulfonic acid into sulphuric and sulfonic
acids by SO2 reduction, according to the following reaction:

RSO4H + SO2 + H2O RSO3H + H2SO4

In this stage two free radicals are generated, therefore no change in the
radical concentration takes place once the stationary state is reached, so that
radicals created during thermal decomposition replace those destroyed by
impurities(X) according to the reaction:

RSO2. + X RSO2X

Irradiated reaction
As can be seen in the upper part of Scheme 2, for each R. produced, one
RSO2. is consumed by collision with the persulfonic acid. Therefore this
stage needs continuous irradiation since radicals that disappear by collision
with impurities have to be replaced by new radicals for the reaction to be
pursued.

Scheme 2.- Proposed reaction mechanism


-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 9

Effect of Wavelength

The optimal conversion-selectivity ratio was obtained at 253.4 nm.


Therefore, all the experiments were conducted at this wavelength.

Effect of Reaction Time

As a general trend, conversion increases with reaction time and


selectivity, expressed as monosulfonate to disulfonate ratio, decreases as
shown in Table 2..
In order to obtain enough sample, the selected time for the batch process
was 6 hours,

Effect of Chain Length

Different fatty acid methyl esters ranging from C12 up to C18, were
sulfoxydized and the reaction conversion obtained for the batch process at 6
hours reaction time.
As mentioned above, after completion of the sulfoxidation reaction, the
reactor product contained mainly mono and polisulfonic acids; some fatty
acids and all the nonreacted fatty acid methyl ester. In order to calculate
conversion, it was necessary first to separate the nonreacted methyl ester from
the reaction products.

Table 2. Φ-MES C16 Conversion-selectivity vs reaction time

Reaction Conversion Mono/di


Time(h) wt/wt % ratio
0.5 4 7.5
1 8 6.2
2 14 4.7
5 33 3.3
6 43 2.3
10 León Cohen David W. Roberts and Claudio Pratesi

Table 3. Gas Chromatographic Analysis of


a C16–C18 Methyl Ester Mixture

Starting material Unreacted material


(mol%) (mol%)
Methyl ester C16 50 59
Methyl ester C18 50 41

Table 4. Average conversion vs chain length

Cn Conversion wt%
C12 24.8
C12 24.4
C12 26.0
C12 26.8
C12 30.0
C12 28.4
C12 Average 26.7
C14 32.0
C14 30.0
C14 35.0
C14 31.5
C14 38.0
C14 38.0
C14 Average 34.1
C16 43.0
C16 42.0
C16 43.5
C16 42.5
C16 42.8
C16 44.0
C16 Average 43.0
C18 50.5
C18 49.8
C18 53.5
C18 48.5
C18 47.8
C18 51.4
C18 Average 50.2
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 11

The conversion was calculated as the recovered non reacted methyl ester
divided by the starting amount. The results of six experiences for each carbon
chain length are shown in Table 4.
The results obtained showed that the conversión increases as the
chainlength increases, going from 30% for C12 up to 50% for C18 [14].
To confirm the relative sulfonatability of the different homologs during
batch sulfoxidation reaction, a commercial 50/50 weight % mixture of C16- C18
was sulfoxidized. A GC analysis was carried out to the methyl ester mixture,
both before and after the reaction. The results (Table 3) indicate that the
relative content of C16 methyl ester increases in the recovered unreacted
material, while the corresponding C18 relative content decreases, meaning that
a higher amount of C18 has been sulfonated as compared to C16.
When two homologs are studied during a sulfoxidation reaction, two
factors have to be taken into account: (i) their molar concentration, which is
related to the probability of collision between both reactants and SO3 (that is,
SO2+ O2), and (ii) their relative reactivity, which is related to the affinity of
each component for SO3 (that is SO2 + O2).
In the sulfoxidation reaction, the main step is radical production.
According to the radical mechanism, radical formation becomes easier the
farther away the CH2 group is from C(O)-OCH3, as a result of the inductive
effect of the latter.
At least, part of the explanation, is that C18 has 15 secondary active
carbons to form radicals as compared to the C16 methyl ester that has 13 sites
(because the primary(ω), the carboxylic and α carbons do not react) . Thus,
the probability of reaction with (SO2 + O2) will be higher for C18.

Analysis

The use of intrumentation techniques such as GC-MS, LC-MS, or IR, in


order to characterize and to identify Φ-MES components, allowed to stablish
for the first time, the composition of reaction products and to demonstrate the
existence of random isomers that were identified [12,13]. The GC-MS analysis
carried out for C12 isomers shown in Figs 1 and 2, indicates that:

a. In Fig.1, two groups of peaks were identified, the first one


corresponds to monosulfonate isomers and the second one to
disulfonate isomers. Monosulfonate amount was 70% (wt/wt) and
12 León Cohen David W. Roberts and Claudio Pratesi

disulfonate 30%(wt/wt). Therefore the selectivity to monosulfonate


was 70%.
b. As seen in Fig.2 and Table 5, no α and a very low amount of β isomer
sulfonate with respect to the carboxylate group, were present. The
distribution is thus random over the δ and ω-1 positions, with a
somewhat lower amount at the γ position. This was due to the fact that
the α does not react and the β carbon reacts very slightly, owing to the
inductive effect of the ester group. The ω carbon is a primary one that
forms a very unstable radical as compared to the secondary ones and
therefore no ω isomer was present. According to the radical
mechanism, radical formation becomes easier the farther away the
CH2 group is from the C(O)-OCH3 as result of the inductive effect of
the latter.

Performance

Surfactants Structure LAS is a mixture of C10 to C13 homologues where


each homologue is a mixture of different isomers with a phenyl ring attached
to the alkyl chain at different positions.

Figure 1. Ф-MES C12 Monosulfonate and disulfonate isomers


-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 13

11
11
6 8 10 11
5 7 5 12
8 6 4 9 1
7 7 4
6 3
5

4
9
8


9

Figure 2. Ф-MESC12 Monosulfonate isomers distribution

Table 5. Ф-MESC12 Monosulfonate isomers distribution.

POSITION DISTRIBUTION (weight%)


3-SULFO(β) 3.2
4-SULFO (γ) 8.4
5-SULFO 12.1
6-SULFO 12.9
7-SULFO 12.7
8-SULFO 12.6
9-SULFO 12.8
10-SULFO 12.8
11-SULFO 12.5
TOTAL 100.0

SAS is a mixture of C14 to C17 homologues, each homologue being a


mixture of different isomers, depending on the position of the SO3 hydrophilic
group.
- MES C16 is a single homologue with different isomers depending on
the position of the SO3 hydrophilic group.
α-MES C16-C18, is a mixture of two linear single chain surfactants (C14*
and C16*), each with a CH(CO2Me)SO3 head group. * We don't count the C
attached to the SO3 group and we don't count the C's of the CO2Me group
14 León Cohen David W. Roberts and Claudio Pratesi

LAS and SAS are both mixtures of homologues with a significantly


higher number of isomers than the single - MES C16 homologue.
-MES C16, SAS and LAS are all double chain surfactants (this is not
strictly a double chain, but a single chain with the polar group attached
somewhere along the chain, thus resulting in branching or ramification or
division in two segments), in each case consisting of mixtures of isomers,
where the lengths of the two chains vary. In SAS and LAS both chains are
alkyl, but in -MESC16 the chains are more different from each other, one
being alkyl and one being terminated by a CO2Me group. The hydrophobicity
of SAS, LAS and -MES C16 is lower than would otherwise be expected
(water solubility higher), because of the water sharing effect between the two
chains, described and modeled quantitatively in [20,21]. The magnitude of the
water sharing effect is different for each isomer. Concerning α-MES C16-C18,
in this case the CO2Me group doesn’t reduce the hydrophobicity of the carbon
chain as in -MES, being within the hydration sphere of the SO3 group which
is at the α carbon.

Water Solubility. Turbidity points, defined as the temperature at which an


aqueous solution of surfactant becomes turbid on cooling, were measured for
the different surfactants as seen in Figure 3. Turbidity points were determined
by cooling, in a thermostated bath, solutions of defined surfactant
concentration until they became cloudy or turbid. For the same surfactant
concentration, the lower the turbidity point temperature the higher the
solubility in water, since the surfactant solution remains clear at lower
temperatures. It must be remembered that turbidity point is the immediate
previous step to crystallization. In Figure 3, the best results correspond to SAS
and LAS, while -MESC16 has acceptable solubilities. All -MES solutions
were cloudy, because these products are not water soluble at laboratory
temperatures (18ºC). Therefore, in this case, turbidity points were measured
after heating solutions until they became clear and then cooling them.
Apparently SAS has a higher water solubility than LAS, and this may be
due to the fact that SAS isomers are more soluble than LAS ones because of
the presence in the latter of the six carbons of the benzene ring.
- MES C16 is less soluble than LAS and SAS, probably because, among
other factors, as it is already well known, a pure homologue is less soluble in
water than a mixture of homologues with the same molecular weight, because
of synergic interactions between homologues and isomers in the mixture. In
order to explain the very low water solubility of α-MES C16-C18 we have to
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 15

consider its structure depicted above, and that the ester group hinders the
sulfonate group and decreases its interaction with water.

Solubility (Turbidity Point)


Temp. (ºC)
30

25

20
Surf. Conc.
15
25%
10 20%
15%
5

0
-MES C16 LAS SAS MES
-5

-10

Figure 3.

STABILITY to WATER HARDNESS

25

20
20 19 19

15
15 14 14
300 ppm
points

11 11 11 11 450 ppm
10
10 600 ppm
8

0
? MES C 16 LAS P-550 SAS α MES C 16-C 18

max. 25 points

Figure 4.
16 León Cohen David W. Roberts and Claudio Pratesi

Stability to water hardness. The determination of the stability to water


hardness ions was carried out through UNE-55-507-72 the Spanish standard
equivalent to the widely used stability test (DIN 53905 or ISO-1063/79). In
the Spanish test, five surfactant solutions with increasing concentrations are
tested for each water hardness level. In this test, a number ranging from 1 to 5
is assigned to surfactant solutions with varying calcium concentrations,
depending on apparent turbidity and precipitate formation: 5 denotes clear
solution, 4 opalescent, 3 turbid, 2 precipitate, 1 precipitate excess. According
to the latter, 25 is the maximum value for each hardness level. As seen in
Figure 4, -MES C16 shows the highest tolerance to water hardness (numbers
are the sum of the five surfactants concentrations tested for each water
hardness level).
Due to the above depicted structure of -MES C16, SAS and LAS
molecules, crystallization of the calcium salt is more inhibited with -
MESC16 than it is with SAS and LAS.
Concerning -MES, since distilled water solutions of -MES are turbid
even at low surfactant concentrations, we cannot measure absolute stability but
relative one, (meaning that, even though the stability number is necessary low,
its value doesn’t change when water hardness increases) and as seen in Figure
4, -MES is not affected by water hardness increase from 150 to 650 ppm.
The reason why the single chain surfactant -MES is reasonably hardness
tolerant (at the higher hardness concentrations it is more tolerant than SAS and
LAS) is because there is some hindrance or shield effect of the CO2ME group
, which results in a weaker interaction with calcium ions, in particular with 2
Ca++. On the other hand, Satzuki [22] claims that LAS and most likely SAS,
are very sensitive to calcium concentration because of the production of a
liquid crystalline calcium salt which is insoluble.  and probably  MES form
a calcium salt that exists in a metastable state, so that might be the reason why
methyl ester sulfonates in general, can be regarded as ones of the most
hardness tolerant surfactants.

Foaming Power. (i) Foam height. The Ross-Miles test was conducted at
49ºC. and 1g/l of surfactant at three different levels of water hardness
expressed as ppm of CaCO3. Experimental results are represented in Figure 5.
-MESC16 gives the best results together with LAS. Water hardness increase
has less effect on both -MESC16 and -MES than on LAS and SAS. This
can be explained based on stability to water hardness of -MESC16 and -
MES (that reaches a maximum at 1g/l and 400 ppm), compared to SAS and
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 17

LAS. (ii) Foam stability. According to the Ross-Miles test, foam stability is
given by the foam height variation after 5 min. The results show that all the
surfactants show similar foam height after 5 min. in all cases than the initial
one.

Initial Foam Height 1 g/l

80
72
70

70
64

60 56 56

53 53

50
150 ppm
mm

40 400 ppm
31 650 ppm
30
23 23
22

20
12

10

0
Φ-MES C 16 LAS P-550 SAS α MES C 16-
C 18

Figure 5.

WETTING POWER 1 g/l

350

300 300 300 300 300 300


300

250 228

200
seconds

150
400
150
650
ppm C aC O3
100

50
50
21
15 12
7

0
Φ-MES C 16 LAS P-550 SAS α MES C 16-
C 18

Figure 6.
18 León Cohen David W. Roberts and Claudio Pratesi

Wetting Power. The Draves test was conducted and 1g/l of surfactant
concentration used at three water hardness levels. According to the
experimental results plotted on Figure 6, -MES C16 is by far the best wetting
agent, specially at higher water hardness. According to Rosen, since the rate of
wetting is a function of the surface tension of the wetting front, the wetting
power of a surfactant is a function of the concentration of molecularly
dispersed material at the front, so we can argue the same argument used for
dishwashing, in the sense that more monomers exist in solution.

Dishwashing. The method used gives a stability index for foam generated
using a certain type of soil, in our case a mixture of olive oil 75% and pig fat
25%, at a definite surfactant concentration [23] and 49ºC. A correlation exists
between the stability index and a hypothetical number of dishes washed
according to the typical dishwashing manual test. A choice has been made so
that the value 50 is an indication when the number of dishes washed is above
50. The results plotted in Figure 7, show that the best performance is reached
with -MES C16 at medium water hardness. This probably reflects the higher
CMC of -MES C16 due to the lower micellar radius (particularly compared
to α-MES), which means that more monomer is in solution to replenish the
foam.

DISHWASHING TEST 0.5 g/l

60

50 50 50 50 50
50

40
Nº Dishes

150 ppm
30 400 ppm
650 ppm
22
ppm CaCO3
20
14
13
12

10
6
5 5

0
Φ MES C16 LAS P-550 SAS α MES C16-C18

Figure 7.
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 19

Skin compatibility: Zein test. The Zein test, a widespread screening in


vitro test for the evaluation of skin irritancy of surfactants, was used. The test
is based on the solubilization of the water-insoluble zein protein by
surfactants. Solubilization of zein is measured through the determination of
nitrogen content of the solubilized protein, giving the so-called Zein number.
According to Kaestner and Frosch [24], the ability of a surfactant to dissolve
the water-insoluble zein protein correlates very well with in vitro test data.
Irritation of human skin is due to the formation of a complex be-tween protein
present in skin and surfactant. According to Seibert and Bolsterdorf [25] the
extent to which a combination of anionic surfactant and protein takes place
depends on many factors. Only surfactant monomers or submicellar species
penetrate membranes, whereas micelles would pre-sumably be too large to
penetrate. Apparently a reduction of critical micelle concentration (CMC)
results in lower levels of free monomers; therefore, surfactants with high CMC
penetrate faster.
The nitrogen content of the solubilized protein gives the so-called Zein
number which classifies anionic surfactants into: <200 mg N/100 mL:
nonirritant; 200–400 mg N/100 mL: moderate irritant; >400 mg N/100 mL:
strong irritant.
The results plotted in Figure 8, show that Φ-MES C16 sodium salt is
nonirritant up to 2.5 wt.% and can be compared favorably to LAS, SAS and -
MES and gives similar results than Alcohol Ether Sulfate [10].

ZEIN TEST

600 564

526

500

400
mgN2/100ml

0,50%
300 288
1%
255
2,50%
200
190 164
169 136
109 108

100 63
85

0
Φ-MES C16 LAS P-550 SAS α MES C16-
C18

Figure 8.
20 León Cohen David W. Roberts and Claudio Pratesi

We see that -MES is much less irritant than LAS and SAS, which can be
explained by its lower to a shield effect of the CO2Me group as mentioned
earlier. Even if Φ-MES C16 has a higher CMC than -MES, its lower Zein
number may reflect weaker hydrophobic binding to the protein, due to the
presence of the ester linkage in the hydrophobe portion of Φ-MES C16. In -
MES the ester linkage is within the hydration sphere of the sulfonate group
and leaves intact the hydrophobic binding of the chain.

REFERENCES
[1] Satsuki, T. (1992). Applications of MES in Detergents, INFORM, 3,
1099-1108.
[2] Satsuki, T., Umehara, K. & Yoneyama, Y. (1992). Performance and
Physicochemical Properties of α -Sulfo Fatty Acid Methyl Esters, J. Am.
Oil Chem. Soc., 69, 672-677.
[3] Steber, J. & P. (1989). Wierich, The Environmental Fate of Fatty Acid
α-Sulfomethyl Esters, Tenside Surfactants Deterg, 26, 406.
[4] Satsuki, T. & H. (1996). Yoshimura,Binding of Calcium Ions to α -
Sulfonated Fatty Acid Methyl Esters in Washing Conditions,4th World
Surfactants Congress, edited by A.E.P.S.A.T., Section D-13, Barcelona,
June 159-168.
[5] Smith, N.R., Alpha-Sulfo Methyl Esters, HAPPI, March: 58(1989).
[6] Dumas et Inchauspé, (1982). Photosulfonation of Normal Paraffins and
Fatty Esters, Adour Enterprise.
[7] Nametkin, N. & Achkasova, L. (1975). Photochemical Sulfoxidation of
Synthetic Fatty Acid Esters, Neftekhimiya, 15, 760.
[8] Beermann, C. (1966). Sulphoxidation of n-Paraffins, European Chem.
News Normal Paraffins Suppl, 2, 37-40.
[9] Asinger, F. (1968). Paraffins Chemistry and Technology, Pergamon
Press.
[10] Cohen, L. & F. (1998). Trujillo, Synthesis , Characterization and Surface
Properties of Sulfoxylated Methyl Esters, J. Surfact. Deterg, 1, 335-341.
[11] Cohen, L. & Trujillo, F. (1999). Performance of Sulfoxylated Fatty Acid
Methyl Esters. J.Surfact.Deterg, 2, 363-365.
[12] Cohen, L., Soto, F., Luna EPSA, M. S., Roberts, D. W., Saul, C. D., Lee,
K. & Williams, E. (2002). UNILEVER; J.E. Bravo PETRESA.
Derivatization, GC-MS, LSIMS and NMR Análisis of Sulfoxylated
-Sulfo Fatty Methyl Ester Sulfonates (-Mes): A New Anionic… 21

Methyl Esters” .. Tenside Surfactants Detergents, Vol.39 nº4. Pag, 78-


83 .
[13] Cohen, L., Soto, F. & Luna, M. S. (2003). from EPSA, C.R. Pratesi, G.
Cassani and L. Faccetti from Sasol Italy. Analysis of Sulfoxylated
Methyl Esters (phi-mes): Sulfonic Acid Composition and Isomer
Identification J.Surfact. Deterg, 6, 151-154.
[14] Cohen, L., Soto, F., Pratesi, C. & Faccetti, L. (2006). (SASOL)
“Sulfoxidation of fatty Methyl esters: Conversión and
Selectivity”.Journal of Surfactants and Detergents, 9, 47-50..
[15] Cohen, L., Soto, F. & Luna, M. S. (2001). Sulfoxylated Methyl Esters as
potential components for liquid formulations. Ibid, 4, 147-150.
[16] Cohen, L., Soto, F. & Luna, M. S. (2001). Separation and Extraction of
-Methyl Ester Sulfoxylate: New Features, Ibid, 4, 73-74.
[17] Cohen, L., Soto, F., Ana Melgarejo, & David, W. (2008). Roberts.
“Performance of -sulfo fatty methyl ester sulfonates (- MES) vs
linear alkylbenzene sulfonates (LAS), secondary alkane sulfonates
(SAS) and -sulfo fatty methyl ester sulfonate (- MES)”. J.Surfact.
Deterg. 11, 181-186.
[18] Cohen, L. & Trujillo, F. “Procedimiento para sintetizar, separar,
caracterizar y analizar fotosulfonatos sódicos a partir de ésteres metílicos
derivados de ácidos grasos de rango detergente “. Boletín Oficial de la
Propiedad Industrial, pag. 9759. 1/12/99.
[19] Cohen, L. & Soto, F. (2006). Patente: “Procedimientos mejorados de
separación y de concentración de productos de la reacción de
fotosulfoxidación de ésteres metílicos grasos de rango detergente.”
Spanish Patent and Trade Mark Office, P20040113Z.
[20] Roberts, D. W. (2000). Aquatic Toxicity - Are Surfactant Properties
Relevant? J. Surfact. Deterg, 3, 309-315.
[21] Roberts, D. W. (2004). Environmental Risk Assessment of Surfactants:
Quantitative Structure-Activity relationships for Aquatic Toxicity, In U
Zoller Ed. Handbook of Detergents: Part B, Environmental Impact.
Dekker, New York. 271-298.
[22] Satsuki, T. et al. (1998). Effect of Calcium ions on Detergency. Tenside
Surfactants Deterg., 35, 112-117.
[23] Soler, J. (1984). Proceedings of the XV Jornadas del Comité Español de
la Detergencia, Barcelona, Spain, 139-169.
[24] Kaestner, W. & Frosch, P. J. (1981). Fette-Seiffen und Anstrichmittel,
83, 33.
[25] Seibert, K. & Bolsterdorf, D. (1988). Magnesium Surfactants - A
22 León Cohen David W. Roberts and Claudio Pratesi

contribution to Mildness, Proceedings of the 2nd Cesio World Surfactant


Congress, Paris, 646-662.

BIOGRAPHIES
Dr. Leon Cohen received his Ph.D. in chemistry at Sevilla University in
1986. In 1994, he earned the EURCHEM designation. He worked for
PETRESA from 1970 to 1996. Since 1989 he is a Professor of Chemical
Engineering at the University of Cadiz, where he is leading since 1993 the
research group entitled "Surface Activity and Renewable Energies". He is the
author of 38 papers and 4 patents related to Detergency and Chemical
Engineering. He is the author of 8 books related to Chemistry and Chemical
Engineering.

Dr. David W Roberts received his Ph.D. in Chemistry from Manchester


University, UK, in 1965. He is a Fellow of the Royal Society of Chemistry and
has the EURCHEM designation. He worked for Unilever Research from 1967
to 2003. Since 2003 he is a consultant in Manufacturing and Toxicological
Chemistry and is an honorary researcher at Liverpool John Moores University.
He is the author of more than 100 papers in the fields of surfactant science and
toxicology.

Dr. Claudio R. Pratesi obtained a degree in biology in 1984 at the


University of Florence (Italy). He then studied chemistry at the Scuola
Normale superiore3 (Pisa) and completed his Ph.D. degree in 1990 with a
study of the conformational structure of dehydropeptides. In 1990 he joined
ENIRICERCHE, the corporate research company of the ENI group (Ente
Nazionale Idrocarburi). He has worked for Sasol Italy (formerly, Condea
Augusta) since 1996.

You might also like