You are on page 1of 8

Biorenewable Resources No.

5 / March 2008

Chemistry of Methyl Ester Sulfonates


David W. Roberts, Liverpool John Moores University, Liverpool, United Kingdom
Luigi Giusti and Alessandro Forcella, Desmet Ballestra S.p.A., Milano, Italy

RCH(CO2Na)SO3Na together with sodium methyl sulfate


Methyl ester sulfonates (MES) are anionic surfactants
(SMS) MeOSO3Na.
with the general structure RCH(CO2Me)SO3Na. They
can be made by sulfonation of saturated fatty acid
2. An aging stage in which the intermediate species react,
methyl esters, RCH2CO2Me derived from natural fats
and the conversion of ME to sulfonated products goes
and oils (1).
to completion. This aging step is much more severe
Interest in MES dates back to at least the early
than in the aging step for linear alkylbenzene (LAB)
1960s, when there were numerous publications by
sulfonation, requiring temperatures of at least 80C. The
Stirton et al. from the U.S. Department of Agriculture
residence time required depends on the temperature, the
and numerous patents from detergent manufacturers.
mole ratio of SO3 to ME, the target conversion level, and
Since that time interest has continued to grow,
the reactor characteristics. Thus, with a batch reactor or
and developments in sulfonation technology have
an ideal plug flow reactor (PFR) and a mole ratio of 1.2,
enabled MES to become an important part of the
45 minutes at 90C or 3.5 minutes at 120C should give
formulators repertoire. Currently there is major
about 98% conversion. With an ideal continuously stirred
interest in MES because of the increasing availability
tank reactor (CSTR), these aging times would need to be
of MES feedstocks with the C 16 derivative (methyl
doubled. Usually this stage is carried out continuously in
hexadecanoate) as the major component, as a by-
reactors whose characteristics are intermediate between
product of biodiesel production (2).
ideal PFR and ideal CSTR.

3. A neutralization stage. If the acidic reaction mixture is


Sulfonation of methyl esters (ME) to produce MES is not neutralized, it deteriorates in color and, particularly
a rather more complex process than sulfonation of other for C16 and higher ME feedstocks, becomes very viscous
major feedstocks. It is now common, with modern reactor and can even solidify unless heated. Neutralization on
technology and present-day feedstock quality, to produce linear a commercial or pilot plant scale is usually carried out
alkylbenzene sulfonates (LAS), primary alcohol sulfates (PAS), continuously in a loop reactor. It is important to avoid
alcohol ethoxysulfates (AES), and alpha olefin sulfonates extremes of pH in neutralization so as to avoid hydrolysis
(AOS) without the need for bleaching (3). In contrast, of MES to di-salt. Neutralization is usually carried out to
ME sulfonation leads to very dark colored products (Klett give a ca. 60% AM paste (AM = Active Matter, in this
values well in excess of 1000) (4). Consequently all current case consisting of MES + di-salt).
commercial ME sulfonation processes require a bleaching The neutralized product from a process involving
step. Other distinguishing features of ME sulfonation are the just these three stages would be a very dark colored
need for a significantly greater than stoichiometric mole ratio paste or solution with a Klett value well above 1000. The
of SO3 to feedstock and the need for a high-temperature aging sulfonation product would consist of a mixture of MES
step. and the di-salt, RCH(CO 2Na)SO 3Na, in proportions of
For MES manufacture at least three stages are essential: ca. 80:20. Sodium methyl sulfate, MeOSO 3Na, is also
present, approximately equimolar with the di-salt. The
1. An ME/SO 3 contacting stage, in which SO 3 is overall reaction is shown in Scheme 1.
chemisorbed by the ME to give intermediate species. If Depending on the formulation in which the MES is to
the mole ratio of SO3 to ME is significantly lower than 1.2, be used, the presence of ca. 20% of the di-salt may or may
full conversion of the ME cannot be achieved. This stage not be acceptable. It is generally regarded as an inferior
is usually carried out continuously in a falling film reactor. surfactant compared with MES. Usually, therefore, two
If the reaction mixture is neutralized at this point, much further steps are involved:
of the ME is recovered unconverted, with conversion
of ME to sulfonated products in the range of 6075%. 4. Because of the high level of color produced, a bleaching
The neutralized sulfonated products at this stage contain step is necessary if the product is to be used for laundry
very little MES, being mainly composed of the di-salt detergents or other consumer products.

www.aocs.org / 2
Chemistry of Methyl Ester Sulfonates

Scheme 1. Overall chemistry of methyl ester sulfonation

5. Depending on the specification required, a re- 1. If this were the only reaction, it should be possible to
esterification step may be included in the process, achieve 100% conversion with an SO3/ME mole ratio of
to convert the di-salt precursor to an MES precursor. 1:1. This is not the case: With 1:1 mole ratio, conversion
This consists of treating the acidic reaction mixture with does not increase beyond about 85% (6).
methanol before neutralization, and this step can reduce
the di-salt content of the neutralized final product to well 2. Kinetic plots show that there are at least two conversion-
below 10% (based on 100% active). increasing reactions taking place in the aging step, with
different rate constants (6).
REACTION CHEMISTRY
Initial reactions and aging reactions Aging kinetics
The initial reaction steps in ester sulfonation occur during the A kinetic model has been developed based on the proposal
ME/SO 3 contacting stage. Probably via complexes formed that two major intermediates are involved in aging (Scheme
reversibly between SO3 and the oxygen atoms of the ester, an 3). One is RCH(SO 3H)COOSO 3Me and the other is a 3:1
-sulfonated intermediate with 2:1 SO3/ME stoichiometry is adduct (or mixture of compounds with overall 3:1 SO 3/ME
formed. This intermediate is believed to have the structure stoichiometry) (6). Although this model is almost certainly
RCH(SO 3H)COOSO 3Me (46). In the aging step, it reacts simpler than reality, it enables kinetic data to be interpreted
with the remaining ME, as shown in Scheme 2. The traditional so as to be able to calculate aging times required for different
interpretation (4) is that this is the only reaction occurring during temperatures and different SO3/ME mole ratios.
aging (excluding color formation, which, despite the intense The concentration of unconverted ME in the reaction
color, consumes only trace amounts of material), and that the mixture decreases with time according to a model (6) based
di-salt and SMS that are found in the neutralized material on two concurrent pseudo first order reactions, one (due to
come from residual RCH(SO3H)COOSO3Me. However, this sulfonation by the 3:1 intermediate) being faster than the other
interpretation is inconsistent with the facts: (sulfonation by the 2:1 intermediate). The rate constants kf and
ks for the faster and slower of these reactions, respectively, can
be calculated from the temperature T (K):

k = A exp (B/T)
for kf (sec1) log A = 12.10, B = 12,060
for ks (sec1) log A = 11.52, B = 12,130

The overall conversion C as a function of time t and mole ratio


M of SO3/ME is given by

C (%) = 100 M (1/M100) 0.25 exp (kst) 0.167 exp (kft)

M100 is the mole ratio that is just sufficient to give a conversion


of 100% after prolonged aging. Rounded to one decimal place,
its value is 1.2. These equations, for aging in a batch reactor
system or in a plug flow system, can be used as guidelines when
setting initial conditions before fine-tuning plant operation to
meet a required specification.

By-products
Significant levels (ca. 5% each) of two other by-products can
be detected in fresh carefully neutralized solutions of MES.
Scheme 2. Traditional interpretation for ME sulfonation These, shown in Scheme 4, are iso-MES, RCH(CO2Na)SO3Me,
stoichiometry and dimethyl sulfoalkanoate (di-MES), which are easily

a special supplement to inform / 3


Biorenewable Resources No. 5 / March 2008

Scheme 3. Model for MES aging kinetics

hydrolyzed to di-salt and MES, respectively (7). Formation methylating species. Dimethyl sulfate could be formed by
of the iso-MES precursor predominates in the early part of attack of ionized MeOSO3H on the methyl group of the mixed
the aging, and the formation of di-MES predominates toward anhydride.
the end of the aging period. Dimethyl sulfate also can be The proposed precursor of iso-MES is the methylated
detected in fresh carefully neutralized solutions of MES, at mixed anhydride (MMA). Since iso-MES is hydrolyzed to
up to 1%, but rapidly decays to undetectable levels (7). If the di-salt, MMA can be regarded as a precursor of the di-salt.
acidic reaction product is treated with methanol before mild The di-acid shown in Scheme 5 is also a precursor of the di-
salt. It is important to note that, since MMA does not have an
ionizable sulfonate group, it cannot undergo the reversible
intramolecular reaction to a cyclic mixed anhydride, which is
proposed as a key step in the release of SO3 during aging (see
Scheme 6). Thus, the SO3 in the form of the OSO3Me group in
MMA is not available as a sulfonating agent. The formation of
MMA explains why a 1:1 mole ratio of SO3/ME is not enough
to give complete conversion. MMA, di-acid, and di-MES are
Scheme 4. By-products of ester all final products in the aging process.
sulfonation
Color formation
As mentioned earlier, severe color formation is a characteristic
neutralization, extra di-MES, but no iso-MES, is observed. of ester sulfonation. ME feedstocks containing unsaturated
This suggests that the iso-MES precursor is very reactive to fatty acid ME yield particularly severe colors, and these have
methanol. been attributed to the formation of poylsulfonated impurities
The simplest interpretation for these by-products is that with conjugated double bonds (8). An unsaturated ME is an
they are formed as a result of disproportionation reactions in internal olefin with a carboxymethyl group at one end of
which the major intermediate, the mixed anhydride, acts as a the hydrocarbon chain. Olefins are very readily sulfonated
methylating agent toward sulfonate groups. In the early stages by SO3, much more so than saturated esters, so if a mixture
of aging, the major species with a sulfonate group is the mixed of saturated and unsaturated ME is sulfonated under MES
anhydride itself, and toward the end of the aging process, the conditions, the unsaturated ester tends to react first, at the
major sulfonate species is MES in its acid form. Scheme 5 double bond, and subsequently oversulfonation and oxidation
represents the overall stoichiometry of the disproportionation of the resulting carboxymethyl internal olefin sulfonate
reactions. It is likely that the detailed mechanisms are more compete with sulfonation of the saturated ester. In practice it
complex than shown, involving dimethyl sulfate as the is very difficult to produce MES of good color quality, even

www.aocs.org / 4
Chemistry of Methyl Ester Sulfonates

Scheme 5. Disproportionation reactions of the mixed anhydride

after bleaching, from an ME with an iodine value (IV) greater opening to a zwitterion, which could lose carbon monoxide to
than 1. Generally, the lower the IV the better, and feedstocks give an alkene sulfonic acid (Scheme 6). Alkene sulfonic acids
hydrogenated to IV values of ca. 0.1 are preferred (1). are formed as major products in alpha olefin sulfonation, and
The foregoing should not be taken to mean that unsaturated alpha olefin sulfonates are very susceptible to color formation
impurities in the feedstock are the only cause of the color if aged in the acid form.
produced in saturated ME sulfonation. Even with laboratory To suppress the unimolecular ring opening of the cyclic
feedstocks having undetectable levels of unsaturation, color -anhydride, and hence to suppress color formation, an
formation is very severe compared with what is experienced extra competing reaction leading to release of SO 3 should be
in, for example, LAB sulfonation. However, the color can provided. Inorganic sulfates should serve this purpose (Scheme
be removed by bleaching much more easily than when the 6) and indeed have been shown to reduce the extent of color
feedstock has IV > 1. formation (10, 11).
The following explanation for color formation from
saturated ME has been proposed (9). The major reaction Bleaching
in the aging step is the conversion of the intermediate All MES processes require a bleaching step. Usually hydrogen
RCH(SO 3H)COOSO 2OMe to RCH(SO 3H)CO 2Me and SO 3, peroxide is used as the bleaching agent, and it can give good
which reacts with more of the ester. The proposed mechanism results when used either before or after neutralization.
is via reversible formation of a cyclic -anhydride and Acid bleaching can be carried out on the acid after the
MeOSO3H, as shown in Scheme 6. As a minor side reaction, re-esterification step, or simultaneously with re-esterification
this -anhydride may undergo reversible unimolecular ring by addition of methanol at the same time. Hydrogen peroxide,

a special supplement to inform / 5


Biorenewable Resources No. 5 / March 2008

Scheme 6. Main aging reaction mechanism, proposed color formation mechanism

usually added at 23%, is used as an aqueous solution (35 or both the sulfonate and the carboxylate groups as their methyl
50%). The water introduced at this stage tends to hydrolyze the esters (Scheme 4), is formed at about 5% level during the
MES acid, which would lead to substantially increased levels acid aging process, and a further 5% is formed from the iso-
of di-salt after neutralization. Residual methanol from the re- MES precursor if methanol is added in a re-esterification step.
esterification, or methanol added at the bleaching stage, can Although di-MES is easily hydrolyzed, the MES neutralization
suppress this hydrolysis and also reduces the viscosity of the step is carried out under mild conditions, to avoid hydrolysis
reaction mixture (1). Without addition of methanol, the di-salt of the MES, and consequently some di-MES can survive.
level would be too high for many applications. Di-MES is electrophilic and reacts with the nucleophilic
Paste bleaching is in many ways simpler than acid hydroperoxide anion, leading to depletion of the bleach. To
bleachingfor example, hydrolysis during bleaching is eliminate this source of inconsistency, the neutralized paste
not a major issuebut has sometimes been considered to can be aged before bleaching (11) to allow the residual di-
be less reliable. This is probably because the implications MES to be hydrolyzed. After paste aging the bleaching step
of the chemistry of the by-product di-MES have not been gives consistently good results that are comparable with acid
fully appreciated. As discussed earlier, di-MES, which has bleaching.

www.aocs.org / 6
Chemistry of Methyl Ester Sulfonates

OCCUPATIONAL AND CONSUMER SAFETY SURFACTANT PROPERTIES AND APPLICATIONS


ISSUES FOR MES OF MES
Like all sulfonate and sulfate surfactants in their acid forms, In Table 1 critical micelle concentrations (CMC) and
MES acid is corrosive. This in itself is no more of a problem Krafft points (T K) are tabulated for MES homologs and the
than with well-established surfactants such as LAS acid. corresponding PAS homologs. For the sodium salts and the
However, there are some special features uniquely associated calcium salts, the CMC values of the two surfactants are
with ester sulfonates. similar when the same carbon numbers (excluding the methyl
Firstly, the methanol injected for re-esterification and/ carbon of the ester group in MES) are compared. However,
or acid bleaching is flammable, having a flash point of 10C the TK values for sodium PAS are much higher (by over 20C)
and being explosive in the range of 544% in air. Appropriate than for the corresponding sodium MES. The difference is
safety procedures for storage and handling therefore need even greater (by over 40C) when TK values of the calcium
to be followed when methanol addition is a part of the MES salts are compared. The di-salt that can result from hydrolysis
process. of MES has a much higher T K value than the corresponding
Secondly, if methanol recovery is part of the process, MES, as illustrated in Table 1 for the C 16 homolog.
precautions must be taken against explosions in the recovery
step, which can happen if peroxides formed in the bleaching Table 1. CMC and TK for MES and linear PASa
step build up. This is potentially more of a problem with acid Surfactant CMC (mmolar) TK (C) Reference
bleaching than with paste bleaching.
Thirdly, the by-products in MES acid are hazardous by Sodium salts
skin contact and, in the case of dimethyl sulfate, by inhalation. C12 MES 5.3 <0 4,12
Dimethyl sulfate, detectable at 1% (100% AM basis) in fresh C12 PAS 8 8 14
carefully neutralized MES solutions, penetrates skin readily C14 MES 2.8 6 4,12
and is a carcinogen in rats. However, it does not survive long C14 PAS 2 30 14
after neutralization and is not a cause for concern regarding C16 MES 0.4 17 4,12
consumer safety. Di-MES, present in MES acid at ca. 510%, C16 PAS 0.40.6 45 14
is a strong skin sensitizer in guinea pigs, but at levels below C18 MES 0.080.16 30 4,12
100 ppm is not considered to give cause for concern on C18 PAS 0.2 56 14
consumer safety grounds. The 100 ppm level may be exceeded C16 Di-salt 65 4
in the freshly neutralized paste, but not in bleached paste or Calcium salts
in dried MES. C14 MES 0.66 28 12
C14 PAS 0.68 71 12
HYDROLYIC STABILITY C16 MES 0.19 41 12
The CO2Me group of MES can undergo hydrolysis to CO2H C16 PAS 85 12
(acid hydrolyis) or CO2Na (alkaline hydrolysis). Thus, MES C18 MES 0.04 49 12
can be hydrolyzed to di-salt or the corresponding acid.
However, due to a combination of steric and electronic
a
Abbreviations: CMC, critical micelle concentration; TK, Krafft
effects of the -sulfo group, the hydrolysis is slower than for temperature; MES, methyl ester sulfonate; PAS, primary alcohol sulfates.
nonsulfonated esters. In the pH range 39.5 the hydrolysis
is very slow (12). Under conditions of acid bleaching with The figures in Table 1 demonstrate that for MES the
aqueous hydrogen peroxide the pH will be below this range, higher homologs, which are more abundant in vegetable oils
which explains why hydrolysis during acid bleaching can be and have better detergency, are sufficiently water soluble to
quite extensive if methanol is not added. In a neutralization be useful in low-temperature laundry products. For PAS the
loop the pH can be controlled below 9.5 so hydrolysis can be higher homologs are too insoluble and the less abundant lower
minimized at that stage. homologs (C 12C 14) are more suitable for low-temperature
Laundry powders have traditionally been made by spray laundry products. It is also clear that MES is much more
drying an aqueous slurry of the surfactants and the inorganic calcium tolerant than PAS. The di-salt, however, is less water
salts in a spray drying tower to form the bulk of the formulated soluble and less calcium tolerant.
powder. Spray drying is still widely used, but nontower routes Various detergency measurements comparing different
are now also used. Under spray-drying conditions MES homologs of MES against each other and against other
undergoes partial hydrolysis to di-salt. Thus, MES is not surfactants have been reported (e.g., 4, 13). Such comparisons do
suitable for formulation by spray drying. However, it can be not always translate well into realistic consumer use situations,
used as the dried solid for nontower production of powders. but the following general conclusions can be drawn:

a special supplement to inform / 7


Biorenewable Resources No. 5 / March 2008

1. The optimal detergency with MES is for the C16 homolog tallow-based MES (C16/C18) fish toxicity EC50 values of 0.40.9
(whose parent ME will be the most abundant if the have been reported (16). However, the primary degradation
feedstock is sourced as a by-product from biodiesel). of ester sulfonates (see below) is fast and would prevent an
accumulation of toxicity.
2. MES detergency is more resistant to water hardness than
detergency of other anionic surfactants. Table 2. Daphnia toxicity of MES
MES EC50a (mg/L)
3. In enzyme detergent formulations, enzyme activity is less
affected by MES than by other surfactants. C12 184
C14 28
4. MES is a good lime-soap dispersant, i.e., when used as
a co-surfactant with soap in hard water it prevents the C16 7
precipitation of calcium soap.
a
EC50, effective concentration at which a 50% response is observed.
These surfactant properties of MES are all consistent with
the concept that the polar but uncharged ester linkage, being The biodegradation characteristics of MES are rather
in close proximity to the negatively charged sulfonate group, similar to those of LAS. Although it is resistant to anaerobic
reduces the charge density of the latter so that electrostatic biodegradation, in aerobic systems MES undergoes rapid
binding to cations is weaker than in simple sulfonates and
biodegradation of the alkyl chain to a slower-degrading
sulfates. A similar effect, in this case the interaction between
residue, which is ultimately completely mineralized (16). The
the ether linkage and the sulfate group, underlies the differences
biodegradation pathway is shown in Scheme 7. The initial
between PAS and AES.
-oxidation step at the end of the alkyl chain is followed by
Because of these properties and because of the perceived
a sequence of -oxidation cycles to arrive at monomethyl
potential for cheap availability of the feedstocks, interest
-sulfosuccinate. This undergoes desulfonation to succinic
has grown in using MES, in combination with other anionic
acid, which features naturally in cell metabolism.
surfactants, in laundry powders and, in markets where soap-
based products are extensively used, in combination with
ESTER SULFONATES OTHER THAN METHYL
soap (4). For these applications a carbon number distribution
So far only ME sulfonates have been developed to the stage of
with C 16 dominant is the optimum. Because of its mildness
production on a manufacturing scale and for use in consumer
to skin and mucous membranes, MES is also of interest for
products. It is conceivable that a future economic situation
dishwashing and shampoo applications, the C12/C 14 carbon
could arise, for example government incentives to combine
numbers being preferred (4). bioethanol technology with biodiesel technology to produce
ethyl ester biodiesel, where ethyl esters become attractive
ENVIRONMENTAL CHARACTERISTICS as sulfonation feedstocks. The sulfonation characteristics of
In aquatic toxicity studies, MES has been shown to behave, ethyl esters are very similar to those of ME, although higher
like other anionic surfactants, as a polar narcotic (15). Table temperatures or longer reaction times are required for the aging
2 shows EC 50 values to Daphnia for C 12 to C 16 homologs. and transesterification stages. It should be quite straightforward
C 18 MES is too insoluble to be tested alone. Intrinsically to adapt an MES production operation to produce ethyl ester
its toxicity is greater than that of the C 16 homolog, and for sulfonates.

Scheme 7. MES biodegradation

www.aocs.org / 8
Chemistry of Methyl Ester Sulfonates

REFERENCES 8. Yamada, K., and S. Matsutani, Analysis of the Dark


1. Roberts, D.W., Manufacture of Anionic Surfactants, in Colored Impurities in Sulfonated Fatty Acid Methyl Ester,
F.D. Gunstone and R.J. Hamilton, eds., Oleochemical J. Am. Oil Chem. Soc. 73:121125, 1996.
Manufacture and Applications, Sheffield Academic Press,
Sheffield, U.K., 2001, pp. 5573. 9. Roberts, D.W., The Origin of Colour Formation in Methyl
Ester Sulphonation, Jorn. Com. Esp. Deterg., 37:153159,
2. Ahmad, S., P. Siwayanan, Z. Abd Murad, H. Abd Aziz, 2007.
and H. Seng Soi, Beyond Biodiesel. Methyl Esters as the
10. United States Patent 6,657,071, to Lion Corporation.
Route for the Production of Surfactants Feedstock, inform
December 2, 2003.
18:216220, 2007.
11. United States Patent Application USSN61/026,174,
3. Roberts, D.W., Sulfonation Technology for Anionic Desmet Ballestra S.p.A., February 8, 2008.
Surfactant Manufacture, Org. Proc. Res. Devel. 2:194202,
1998. 12. Stein, W., and H. Baumann, -Sulfonated Fatty Acids
and Esters: Manufacturing Process, Properties, and
4. Schwuger, M.J., and H. Lewandowski, -Sulfomono- Applications, J. Am. Oil Chem. Soc. 52:323329, 1975.
carboxylic Esters, in H.W. Stache, ed., Anionic Surfactants,
Organic Chemistry, Vol. 56 in Surfactant Science Series, 13. Satsuki, T., Applications of MES in Detergents, inform 3:
Marcel Dekker, New York, 1995, pp. 461500. 10991108, 1992.

5. Schmid, K., H. Baumann, W. Stein, and H. Dolhaine, 14. Domingo, X., Alcohol and Alcohol Ether Sulfates, in H.W.
Proc. 1st World Surfactants Congress, Munich, Vol. II, Stache, ed., Anionic Surfactants, Organic Chemistry, Vol.
105, 1984. 56 in Surfactant Science Series, Marcel Dekker, New
York, 1995, pp. 223312.
6. Roberts, D.W., P.S. Jackson, C.D. Saul, C.J. Clemett, and 15. Roberts, D.W., S.J. Marshall, and G. Hodges, Quantitative
K. Jones, A Kinetic and Mechanistic Investigation of Ester Structure-Activity Relationships for Acute Aquatic
Sulphonation, Proc. 2nd World Surfactants Congress, Toxicity of Surfactants, World Surfact. Congr., 4th, 4:
Paris, Vol. II, 3841, 1988. 340351, 1996.

7. Roberts, D.W., C.J. Clemett, C.D. Saul, A. Allan, and 16. Gode, P., W. Guhl, and J. Steber, kologische Bewurtung
R.A. Hodge, Intermediate By-products in Methyl Ester von -Sulfofettsauremethylestern, Fat Sci. Technol.
Sulphonation, Jorn. Com. Esp. Deterg. 26:2733, 1995. 89:548552, 1987.

a special supplement to inform / 9

You might also like