Está en la página 1de 13

Topics in Catalysis Vol. 27, Nos.

14, February 2004 (# 2004)

105

Solid catalysts for the synthesis of fatty esters of glycerol,


polyglycerols and sorbitol from renewable resources
Carlos Marquez-Alvarez, Enrique Sastre, and Joaqu n Perez-Pariente
Instituto de Catalisis y Petroleoqumica, CSIC. C/Marie Curie, S/N. Cantoblanco, 28049-Madrid, Spain

Partial esters of polyols such as glycerol, polyglycerols and sorbitol are nonionic surfactants used as emulsifiers in the food,
detergents and cosmetics industries. The commercial large-scale production of these chemicals relies on traditional homogeneous
catalysis processes making use of strong mineral acids or alkalis. This technology possesses severe drawbacks, related to the
generation of large amounts of by-products, high energy demand, and heterogeneity of the ester mixtures obtained. The
development of new processes based on more selective solid catalysts has, therefore, a great economical interest. This contribution
reviews the scientific and technical literature in this field. It is proposed that solid catalysts based on MCM-41 and other
mesostructures could give rise to the development of new, more efficient processes for the large-scale synthesis of fatty acid esters of
polyols.
KEY WORDS: fatty acid esters; polyols; monoglycerides; sorbitol; polyglycerol; nonionic surfactants; emulsifiers; heterogeneous
catalysts; zeolites; mesoporous catalysts; esterification; transesterification; glycerolysis.

1. Introduction
There is an increasing trend in the chemical industry
to introduce new processes that should meet requirements such as generation of nearly zero waste chemicals,
reduced energy input and use, whenever it is possible, of
raw materials derived from nonfossil primary sources.
Besides, a higher efficiency is generally required, in
terms of optimum transformation of the feedstocks into
the desired product. Many of these requirements
challenge conventional processes based upon old technology, and therefore act as a strong driving force for
innovation, which includes the use of heterogeneous
catalysts.
Despite the enormous impact that solid catalysts have
on the production of chemicals, ranging from large-scale
refining and petrochemical processes to fine and
specialty compounds, their potential is far from being
fully developed in some key areas of the chemical
industry. A notorious example of the prevalence of aged
technology based on homogeneous catalysts is provided
by the commercial production of fatty acid esters of
polyols. Among them, those of glycerol, polyglycerols
and sorbitol will be most relevant to this review.
Fatty acid esters of these polyols, particularly those
of glycerol (about 80% of the total), are widely used as
emulsifiers in a large variety of applications, particularly
in food processing. The hydrophobic acyl group of the
ester is formed by a hydrocarbon chain containing 10 to
20 carbon atoms, whereas the polar head of the
surfactant is formed by hydroxyl groups, the number of
 To whom correspondence should be addressed.
E-mail: jperez@icp.csic.es

which varies from 2 for monoglycerides to around 5 for


some polyglycerol esters (PGE).
The emulsifying properties of the esters are basically
dependent upon their hydrophiliclipophilic balance,
best known as HLB [1]. According to Griffins scale, the
higher the HLB value, the greater is the hydrophilicity
of the surfactant. A large range of HLB values can be
covered by proper combination of the number and
nature of the hydrophobic hydrocarbon tail and the
number of free hydroxyl groups contained in the polar
head of the ester. Therefore, it is possible to select
specific surfactants for targeted applications as a
function of their HLB values. For practical purposes,
and owing to the strong hydrophobic character of the
fatty acyl group, esters containing an OH/fatty chain
ratio equal or higher than two are, in general, best
preferred as emulsifiers. Thus, in the case of glycerol,
only the monoesters are good emulsifying agents. The
same applies to sorbitol, despite this polyol contains six
hydroxyl groups. This is so because, due to its low
thermal stability, sorbitol experiences anhydrization at
the temperature at which commercial esterification is
currently carried out (about 150  C or higher). Under
these conditions, sorbitol transforms into a mixture of
inner ethers, sorbitans and sorbides, which possess 4 and
2 free hydroxyl groups respectively. Therefore, only the
so-called sorbitan monoesters exhibit suitable emulsifying properties.
Details on polyglycerol chemistry will be given below,
but it suffices to say here that the number of hydroxyl
groups is given by n 2, n being the number of glycerol
units making the polyglycerol molecule. For lower
polyglycerols, monoesters are preferred, whereas diesters and eventually even triesters of polyglycerols of
1022-5528/04/02000105/0 # 2004 Plenum Publishing Corporation

106

C. Marquez-Alvarez et al./Fatty acid esters of polyols

Figure 1. Scientific and technological publications on the synthesis, processing and use of (a) fatty acid monoglycerides, (b) polyglycerol esters
and (c) sorbitol monoesters. Insets: distribution of patent applicants nationalities. Data collected using the Chemical Abstracts Service research
tool SciFinder1. *Data on 2003 corresponds to the period from January to April.

high molecular weight could still have good emulsifying


properties.
Figure 1 displays the evolution of the number of
publications on monoglycerides, polyglycerol esters and
sorbitol monoesters, covering application, synthesis and
processing, both in journals and patents. Noticeable
differences in scientific and technological activities
among the three groups of esters are observed.

Significant research effort in monoglyceride chemistry


started in the thirties of the last century, and has been
increasing steadily since then. A relatively large number
of journal publications have been continuously released,
although the relative number compared to patents
started to decrease in recent years. Research on fatty
esters of polyglycerol showed a significant level in the
sixties and seventies, but it started to grow in the early

C. Marquez-Alvarez et al./Fatty acid esters of polyols

eighties at a much faster rate than that of monoglycerides. It is remarkable that this intense research activity
has given rise to a large number of patents, while very
few contributions in this field have been published in
scientific journals. The scientific/technological activities
concerning sorbitol esters have grown at a much lower
rate than those of the two other polyol esters, and very
few journal reports have been published so far on
sorbitol monoesters. This fact might reflect the relative
commercial impact of these esters.
The statistics given in figure 1 clearly show strong
and increasing innovative activities on PGE, which
nevertheless remain virtually unexplored outside the
industrial R & D departments. An analysis of the
nationalities of the patent applicants (figure 1(b), inset)
reveals the leadership of Japanese companies, particularly in the expanding PGE market.
The commercial synthesis of the fatty acid esters of
polyols is carried out basically by two different
procedures: direct esterification of the polyol, usually
catalyzed by a homogeneous acid, such as sulfuric and
sulfonic acids, or transesterification of triglycerides and
polyols, catalyzed by alkaline hydroxides like NaOH,
KOH, CaOH2 , or the sodium salts of lower aliphatic
alcohols, like methanol. It can be estimated that
approximately 80% of the manufacturing processes
belong to the first pathway.
The fatty acids and the glycerol are obtained basically
by splitting of fats and vegetable oils with water at high
temperature [2]. Sorbitol is also derived from nonfossil
sources, as it is obtained by catalytic hydrogenation of
glucose, starch being the basic commercial source of this
sugar. Therefore, a variety of abundant raw materials
resulting from agriculture activities are readily available
for the synthesis of the polyol esters described in this
work.
Here, we provide an overview of the synthesis
procedures of these polyol esters and discuss the
potential that exists for developing new, more efficient
synthesis processes based on solid catalysts. Enzyme
catalyzed synthesis of polyol esters is not covered in this
review. An intense research effort is being carried out on
the use of lipases to synthesize esters of glycerol, sorbitol
and sugars [36] and even polyglycerol [7,8]. These
studies show that enzymes have an enormous potential
as catalysts in those processes in which high regioselectivity is required. However, for large-scale synthesis of
the more common surfactants, these processes are not
yet competitive due to the high cost of enzymes and their
low productivity.

2. Monoglycerides
The product of the direct esterification or transesterification of glycerol is a mixture of mono-, di- and some
triglyceride, and dissolved glycerol, which typically

OOC R
OOC R

107

OOC R

OH

OOC R

OH
OH

OH
OH

Scheme 1.

contains 4060% monoglyceride and 3545% diglyceride. Most of the fatty monoglycerides are obtained by
transesterification of triglycerides, either from vegetable
or animal sources, and glycerol. The basic characteristics of the process have been comprehensively reviewed
[9]. Scheme 1 describes the overall simplified chemical
equation for glycerolysis of fats and oils.
The actual transesterification process is more complex than that described by scheme 1, in such a way that
di- and even triester are simultaneously obtained.
Typical composition of the esters blends has been given
above. Besides, scheme 1 does not reflect the formation
of the 2-acyl or -monoglyceride, the proportion of
which remains below 10% at low temperature but
increases with the reaction temperature.
The basic catalysts used in the process, generally
KOH and CaOH2 , are neutralized with phosphoric
acid after the reaction, and the formed salts are
removed. Therefore, undesired waste chemicals are
obtained. The use of solid catalysts suitable for
regeneration would obviously avoid these drawbacks.
However, this is not the only driving force to modify the
classical process. The glycerolysis reaction requires a
high temperature in order to increase the solubility of
glycerol in the fat phase. An additional energy-consuming step is needed to separate the monoglyceride from
the crude reaction product by vacuum molecular
distillation in order to obtain the monoester with a
purity of 90% at least. Therefore, the key aspect of the
process is the low selectivity to the monoglyceride. It
must be stressed that direct esterification of the glycerol
with a fatty acid, by using homogeneous acid catalysts,
does not improve the monoglyceride selectivity.
These considerations have led to explore alternative
synthesis routes that involve in most cases a change of
the nature of the catalyst and the chemical feedstocks.
The use of inorganic solid catalysts will now be
discussed.

2.1. Base catalysis


Solid basic materials such as MgO, AlMg hydrotalcites as well as Cs-exchanged sepiolite and mesoporous MCM-41 have been used as catalysts for
the glycerol transesterification with triglycerides [10].
MgO and Al-poor hydrotalcites Al=Al Mg 0:20
behave as active catalysts for triolein glycerolysis, but a
glycerol/fat molar ratio of as high as 12 is required to
obtain a monoglyceride yield of 70% at 97% conversion
(MgO catalysts at 240 8C). The large excess of glycerol

C. Marquez-Alvarez et al./Fatty acid esters of polyols

108

requires an increase of the reactor capacity of over 60%,


compared to stoichiometric reaction mixtures, and
increases the thermal demand by a factor of nearly 2.
This example illustrates the negative influence of the low
solubility of the hydrophilic glycerol in the fat phase.
The use of organic solvents would have benefits on the
overall reaction but, for safety reasons, it is avoided
when producing monoglycerides to be used as food
additives.
Other esters have been used for glycerol transesterification. Metal oxides like MgO, CeO2 , La2 O3 and ZnO
have been used as solid base catalysts for the transesterification of glycerol with stoichiometric amounts of
methyl stearate [11] in the absence of solvent. The
catalysts are active, but the selectivity to mono-, di- and
triesters is similar to that obtained by using homogeneous basic catalysts (40% monoester at 80% conversion).

2.2. Acid catalysis


A variety of solid acids have been used to catalyze the
esterification of glycerol with fatty acids. Zeolite
materials meet, in principle, the basic requirements for
successful replacement of homogeneous acids, as they
possess sufficient acid strength. Indeed, they have been
used for the esterification of monoalcohols with low
molecular weight carboxylic acids [12]. When applied to
glycerol esterification with fatty acids, the relatively
small pore aperture of zeolite materials would have the
potential to reduce the formation of the bulky di- and
triesters, increasing therefore the selectivity to monoglycerides. Table 1 summarizes the most relevant results
reported when zeolite catalysts are used for the
esterification of glycerol with lauric and oleic acids.
Large-pore tridirectional 12-membered-ring structures, like faujasite and beta, are more active and

selective than 10-membered-ring structures, like ZSM-5,


or unidirectional 12-MR materials like mordenite [13
18]. Influence of the aluminum content of the framework has been reported for faujasite [14] and beta [18].
For both structures, the conversion and monoglyceride
selectivity increase with the Si/Al ratio, and a maximum
is observed at Si/Al ratios in the range of 3540.
Absence of crystal size effect for dealuminated faujasite
Si=Al 14 in the range of 0:33:5 m has been taken
as a proof that the esterification with lauric acid takes
place inside the intracrystalline voids of the zeolite [16].
In this way, transition state selectivity inside the
supercages suppresses the consecutive esterification of
the monoglycerides to bulkier glycerides. A similar effect
would explain the high selectivity of zeolite beta.
However, as di- and even traces of triglycerides are
readily observed even at conversion higher than 10%, a
significant contribution of the external surface to the
overall reaction cannot be totally excluded. Indeed, both
the activity and the selectivity with oleic acid are much
lower (table 1), which suggests a severe pore-size
limitation to the reaction rate.
In order to avoid the pore-size constraints of the
zeolite materials on the reaction, cross-linked porous
polymers, like ion-exchange resins containing sulfonic
acid groups, have been used as catalysts for the
esterification of glycerol with fatty acids [19]. It has
been observed that gel resins, like Amberlyst, are more
active than macroporous resins owing to their swelling
capacity. Both the activity and selectivity show a strong
dependency on the glycerol/fatty acid ratio (R), as well
as on the chain length of the fatty acid. The reaction is
faster with lauric acid than with oleic acid. For lauric
acid, using Amberlyst 31 as catalyst, the activity
decreases with R, but the monolaurate selectivity
increases up to R 3. For R 6 and 140  C, the
selectivity to monolaurate is 83% at 65% conversion.

Table 1
Esterification of glycerol with fatty acids catalyzed by zeolites
Catalyst

Si/Al
ratio

Fatty
acid

Glycerol to
fatty acid
molar ratio

Reaction
temperature
(8C)

Fatty acid
conversiona
(%)

Monoglyceride
selectivitya
(%)

References

FAU
FAU
FAU
FAU
FAU
BEA
MFI
FAU
MOR
BEA

10.0
37.5
7.3
7.7
14.8
14.0
16.5
35.0
62.0
42.0

Oleic
Oleic
Oleic
Oleic
Lauric
Lauric
Lauric
Lauric
Lauric
Lauric

6
1
1
1
1
1
1
1
1
1

90
100
100
180
112
112
112
100
100
100

5
10.5
6.9
80b
45
38
33
13
16
32d

67
53
53
76b
69c
64c
57c
68
55
68d

[13]
[14]
[14]
[15]
[16]
[16]
[16]
[17]
[17]
[18]

Reaction time 24 h.
Reaction time 5 h.
c
At 50% conversion.
d
Reaction time 10 h.
b

C. Marquez-Alvarez et al./Fatty acid esters of polyols

The selectivity for oleic acid at the same conversion level


is lower. However, the selectivity is higher than that
obtained with zeolite materials. No information on
catalyst regeneration is provided in these works.

2.3. Acidic mesoporous materials


The benefits of the swelling of the ion-exchange resins
on the overall catalyst performance in glycerol esterification evidence the positive influence of increasing the
accessibility of the reagents to the active centers.
Therefore, the family of the ordered mesoporous
materials appears as a very interesting alternative for
the development of solid catalysts, as it offers the
possibility to overcome the pore-size limitation characteristic of zeolite materials. The pore size of these
materials is typically in the range of 210 nm, and
several structures possessing different pore dimensions
and channel and voids architecture are available.
Unfortunately, the strength of the acid sites present in
conventional MCM-41 materials is low [20], and,
consequently, their activity and monoglycerides selectivity are low [14].
As the sulfonic acid groups present in the ionexchange resins are very active in catalyzing the
esterification reaction, they have been attached to pure
silica mesoporous frameworks in order to obtain
mesoporous catalysts. The behavior of mesoporous
materials containing R-SO3 H groups in the esterification of glycerol with fatty acids has been reviewed
recently [21], and some of the most relevant aspects are
briefly commented here. MCM-41 materials containing
propylsulfonic acid groups have been reported to
catalyze selectively the esterification of glycerol with
lauric acid for equimolar amounts of both reagents, and
in the absence of solvent. A monolaurin yield of 53%
was obtained, whereas for oleic acid the monoglyceride
yield dropped to 31% [22]. Following this pioneering
work, further improvement of the SO3 H-containing
MCM-41 catalysts prepared by a one-pot synthesis
procedure was obtained through the optimization of the
pore arrangement by using mixtures of surfactants
having different chain lengths [23], as well as mixtures
of n-dodecylamine and hexadecyltrimethylammonium
cations in the synthesis gel [24]. While these modifications increase the activity and monoglyceride yield of the
reaction, a major improvement in catalyst efficiency
results by tuning the hydrophilic/lipophilic balance of
the catalyst surface [25]. The anchoring of hydrocarbon
moieties like methyl and propyl groups decreases
substantially the strong affinity of the catalyst to
hydrophilic molecules, like water and glycerol. As a
consequence, the monolaurin yield raises to 66%, while
that of monoolein is close to 50% [26]. The catalysts can
be reused, although they show a small loss of activity
[23]. Sulfur leaching is small. After 8 h at 100  C, the

109

sulfur content in the reaction mixture is lower than


80 ppm, although it increases with the temperature and
reaction time.
Organosulfonic groups other than propylsulfonic
have also been attached to the MCM-41 surface by
one-pot synthesis procedures. For example, the intrinsic
activity of the phenylsulfonic acid-containing material is
higher than that with propylsulfonic acid, although the
selectivity is lower [27].
MCM-41 catalysts have also been prepared from gels
containing 3-mercaptopropyl(methyl)dimethoxysilane.
The presence of a methyl group attached to the silicon
atom bonded to the propylsulfonic moiety increases
substantially both the activity and selectivity. At 50%
conversion, the selectivities to monolaurin and monoolein were 77% and 65% respectively [28].
More recent results evidence the influence of the nalkyl chain length of the organosulfonic acid moiety in
glycerol esterification [29]. Catalysts obtained by sulfonation of MCM-41 materials containing vinyl groups
exhibit better activity and selectivity than propylsulfonic
acid-containing materials.
The influence of the pore size of functionalized
MCM-41 catalysts containing sulfonic acid groups has
not yet been properly addressed. Catalysts having lower
activity and selectivity are achieved by replacing methyl
groups by propyl groups in catalysts containing both nalkyl and propylsulfonic moieties [26]. When performing
studies on the influence of pore size in functionalized
materials, it has to be remarked that the effective pore
size varies with the S-loading and with the size of the
functional group [30].
Comparison of the results obtained from lauric and
oleic acids suggests that the optimum pore size for the
reaction varies according to the chain length of the fatty
acid [26]. Indeed, it has been shown recently that the
selectivity to glycerol monostearate of MCM-41 functionalized materials having a S-loading of 0:3 meq g1
decreases from 47 to 40% at 50% acid conversion, as the
pore size increases from 2.6 to 3.7 nm [31]. A further
decrease of the pore size to 2.0 nm does not change the
selectivity, but decreases the turnover values by 80%.
These results suggest that there is a noticeable contribution of the external surface to the overall reaction for
small pore materials.
Mesopore structures other than MCM-41 have been
explored as catalysts for glycerol esterification. SBA-15,
which possesses unidirectional channels arranged in
hexagonal symmetry and interconnected by micropores,
has been functionalized with propylsulfonic acid groups
0:78 meq g1 . Its specific activity (TON) and monoolein selectivity are lower than that of MCM-41
catalysts, which has been attributed to both the presence
of nonaccessible micropores, and to the large mesopore
size, reaching 3.2 nm [32]. SBA-12, a structure with
interconnected spherical cavities in a cubic-close packing
[33], and SBA-2, an intergrowth of hexagonal and cubic

C. Marquez-Alvarez et al./Fatty acid esters of polyols

110

mesostructures, both of which contain a complex set of


narrow channels that interconnect the large cavities,
show low activity and selectivity [32,34]. Therefore, none
of these complex structures afford better catalysts than
conventional MCM-41.

OOC R

R COOH

OH

OH

OH
Scheme 2.

3. Polyglycerol esters

2.4. Basic MCM-41 catalysts


Despite the improvement in monoglyceride selectivity
achieved by using mesoporous acid catalysts, a monoglyceride yield of 90%, that would be needed in order to
avoid the costly molecular distillation of the ester
mixtures, has not yet been reached.
Nucleophilic ring opening of glycidol with fatty acids
(scheme 2) has been explored as a selective and
convenient route toward nearly pure monoglycerides
[35]. To accomplish this purpose, MCM-41 all-silica
materials have been functionalized by grafting R-propyl
moieties, where R stands for NH2 and a variety of cyclic
secondary amines. In this synthesis route, capping of the
residual silanol groups with trimethylsilyl groups, for
example, is required to prevent glycidol polymerization.
However, reused catalysts also show improved activity
because of the grafting of the residual silanol groups
with glycerol molecules during the first run. Catalysts
containing tertiary and hindered amines are more
efficient than those having primary amines owing to
steric effects [36]. Indeed, the activity correlates
approximately with the base strength of the organic
base. Following this procedure, monolaurate yields of
90% have been reported. This synthesis route would be
attractive for monoglyceride production, were it not for
two reasons: (1) the glycidol is much more expensive
than glycerol, the price of which goes down as its
production increases because of the increasing use of
biodiesel, and (2) toluene is used as solvent in the
addition reaction of glycidol to fatty acid. Both factors
make this route unattractive for commercial monoglyceride production.

As it has been described above, the emulsifying


properties of the PGE depend basically upon the length
of the polyglycerol chain, the degree of esterification and
the molecular weight of the fatty acid. Indeed, the
preparation of PGE, suitable as additives in foodstuffs,
is a process much more complex than that of monoglycerides, and, in general, involves two different steps:
(1) polymerization of glycerol in the presence of a small
amount of alkali (base catalysis), and (2) esterification
of the resulting polyglycerol. Scheme 3 offers an
overview of the process.

3.1. Polyglycerol synthesis


Glycerol polymerization is carried out at a temperature above 200  C under alkaline conditions. The
condensation reaction involves the terminal hydroxyl
groups of glycerol molecules as they are more reactive
than the 2-hydroxy group. However, some ramified and
even cyclic by-products are always formed. Carbon
dioxide or nitrogen is bubbled through the reaction
mixture to prevent dehydration of glycerol to undesired
acrolein. Several bases have been tested as catalysts,
including hydroxides, carbonates and oxides of several
metals [37]. It has been found that the carbonates are
more active than hydroxides, despite the weaker base
character of the former. This has been attributed to the
better solubility of carbonates in the glycerol and in the
polymeric product at elevated temperatures. Oxides like
MgO, CaO and ZnO are less active because of lack of
solubility.
The polymerization reaction proceeds by chance
under homogeneous conditions and therefore a mixture
of products with different molecular size is obtained [38
44]. Low amount of catalyst <1 mol% and relatively
OH

OOC R

OH

OH

O
OH
(n+2)

OH

- (n+1) H2O

O
R-COOH

OH

OH

- H2O

OH
O

O
n

OH

OH

OH

OH

Scheme 3.

C. Marquez-Alvarez et al./Fatty acid esters of polyols

3.2. Esterification of polyglycerols


The esterification of polyglycerols leads to mixtures
with a broad range of different compounds, owing to the
large number of hydroxyl groups and the intrinsic
complexity of the commercially available polyglycerols
mixtures. The hydrophiliclipophilic properties of the
esters, expressed by means of the HLB concept [1], are a
function of the molecular weight of the esterified fatty
acid, the degree of esterification of hydroxyl groups and
the number of glycerol units forming the polyglycerol
backbone. As an example, figure 2 shows the range of
HLB values of triglycerol esters as a function of the
carboxylic acid chain length and the degree of esterification. It can be concluded that a decrease of the
molecular weight of the fatty acid results in an increase
in the hydrophilic nature of the ester (the HLB value
increases). On the other hand, an increase in the number
of fatty acid moieties that react with the polyol molecule
will result in a more lipophilic product. For a given fatty
acid and esterification degree, the hydrophilic character
increases with the molecular weight of the polyglycerol
(higher hydroxyl/fatty acid ratio). This is depicted in
figure 3 for stearic acid esters of several polyglycerols.
As shown in figure 3, an HLB range from about 3 to 8
can be covered by different polyesters of tri- to
decaglycerol, but monoesters of high molecular weight
polyols are required to get highly hydrophilic compounds (HLB  12).
Figure 3 underlines the strong impact that the chemical
composition of the polyglycerol has on the emulsifying
properties of the esters resulting from the esterification
process. A narrow distribution of the molecular weight of
polyglycerols would be needed in order to obtain products
having well-defined HLB values.
16

14

12

Esterification
degree

10

HLB

short reaction time are required to exert some control on


the polymerization degree in order to favor a mixture of
products having, on average, low polymerization degree.
Product distribution of polyol is important as the
polyglycerol moiety of the ester should meet some
specifications to be used as food additive. According to
EC regulations, the content of di-, tri- and tetraglycerol
should predominate, whereas the content of polyglycerols equal to or higher than heptaglycerol must be low [45].
Food and Drug Administration regulations allow the use
of polyglycerols up to and including decaglycerol [46].
The alkaline polymerization of glycidol affords a
more selective process, provided that a strict control of
the reaction parameters is exerted [47]. As an example,
linear octaglycerol has been prepared from glycidol by
using KOH as catalyst at a reaction temperature of
ca. 120  C [48]. The secondary hydroxyl content was
72% (theoretical maximum 80%). In contrast, the
octaglycerol obtained by conventional glycerol condensation at 230250  C, using NaOH as catalyst, contained
41% secondary hydroxyls, due to the presence of
residual glycerol and oligomers of lower molecular
weight. Besides, the condensation reaction of glycerol
and glycidol catalyzed by acids has been claimed to
render polyglycerols of high polymerization degree with
low color values [4951].
Porous solid catalysts could exert some shapeselective effect on the course of the polymerization
reaction. Nevertheless, reports on such catalysts are
extremely scarce. Zeolites NaX and NaA have been used
as catalysts for the selective production of diglycerol
[52]. Cs-exchanged zeolite X shows high selectivity to di(62%) and triglycerol (33%), at 70% conversion, while
only a 4% of tetraglycerol is obtained [53,54]. On the
other hand, medium-pore Cs-containing zeolites like
ZSM-5 are less active and selective.
The influence of pore size on the polymerization
reaction is evidenced by the use of basic MCM-41
catalysts [55]. The Cs-impregnated material affords the
best results, whereas Mg- and La-containing catalysts
favored glycerol dehydration and the formation of
acrolein. As an example, for a conversion of 80%, the
selectivity to di- triglycerol (di-/triglycerol > 1) is
close to 90%. However, a severe leaching of cesium is
observed, and a decrease of the cesium loading of the
catalysts between 95 and 70% has been reported. Indeed,
no X-ray diffraction peak at low angle is observed
for cesium loading equal to or higher than
18:6  104 eq g1 . Despite the severe cesium leaching,
the catalysts still show a remarkable high selectivity to
di- triglycerol.
These results are promising, but, nevertheless, the
selectivity of these mesoporous catalysts approaches
that of the Cs-exchanged faujasite X. They also evidence
that there is much need for investigation into reusable,
shape-selective catalysts for controlled glycerol polymerization.

111

2
4

3
4
5

0
2

10

12

14

16

18

Fatty acid chain length (No. of C atoms)


Figure 2. HLB values of triglycerol fatty acid esters.

20

C. Marquez-Alvarez et al./Fatty acid esters of polyols

112

[40,56,6668]. The reaction can also be carried out by


transesterification of a polyol with a triglyceride or a
fatty acid alkyl ester (methyl ester usually), generally in
the presence of a suitable alkaline catalyst [40,4244,69].
Furthermore, it has been reported that partial esters of
polyglycerol can be obtained by acid-catalyzed polycondensation of monoglyceride or monoglyceride and
glycerol [70,71]. Finally, the addition polymerization of
glycidol to a fatty acid or to a fatty acid monoglyceride
catalyzed by acids also renders PGE [7277]. These
reactions must be carried out under an inert atmosphere
to obtain good color and organoleptic properties. Some
examples describe the use of moderate vacuum as well.
In the commercial manufacture of partial esters, the
reaction will give a mixture in which the fatty acid
radicals are distributed among all available hydroxyl
groups, whose proportions are dictated basically by the
thermodynamic constraints of the reaction, namely,
temperature and molar ratio of the reagents. As an
example, the product obtained by reacting equimolar
quantities of triglycerol and stearic acid contains 41 and
31 wt% of mono- and distearate respectively [41]. As a
consequence, the HLB value of the product is lower
than that of the pure monoester. To approach the
theoretical HLB values of the monoesters, a large excess
of polyglycerol is required.
Owing to the inherent complexity of the esters
mixtures that are obtained by these methods, in most
of the publications these mixtures are characterized in

Polymerization
degree
10

16

14

7
12

5
4

HLB

10

3
8

0
0

10

Esterification degree
Figure 3. HLB values of polyglycerol stearic acid esters.

The processes for manufacturing PGE are summarized in scheme 4. The esters are most simply made by
direct esterification of the polyol with a free organic
acid. The reaction can proceed at T > 200  C without
using a catalyst [37,5658], or at T < 200  C by
using either alkali [37,56,5965] or acid catalysts

Direct esterification
OH OH

OH

OH OH

OH OH

OH

RCOOH

OH OOC-R

H2O

Transesterification
OH OH

OH

OOC R

OH OH

OH

OH OOC-R

OH OH
RCOOMe

OH
O

OH
OOC R
OOC R

OH OH

OH

OOC R

OH OH

OH OH

OOC R

OH OOC-R

MeOH

Polycondensation
OH

OH

OH

(n+1)

OH OH

OH

OH

OH OOC-R

OH

OOC R

+ (n+1) H2O

O
n

Addition polymerisation of glycidol


O

OH

RCOOR'

- R'OH O

OOC-R (n+1)

OH

H2O

OH OH

OH
O

OH OOC-R
O
n

OH OH
R' = H , CH2 CH CH2

Scheme 4.

C. Marquez-Alvarez et al./Fatty acid esters of polyols

3
4
5

-1

Saponification No. (mg KOHg )

200

150

2
43
95 3
8 4
7

es
ter
i

f ic
ati

on

100

glycerol

3
5

de

diglycerol

2
1 triglycerol

gr e

pentaglycerol

e
2

decaglycerol
50
0

100

200

300

400

500

600

-1

OH No. (mg KOHg )

Figure 4. Saponification and OH numbers of oleic acid esters of


polyglycerols. Figures indicate the esterification degree (number of
fatty acyl groups per molecule) of each ester.

terms of their saponification and OH values. These


numbers are related to the mean polyglycerol chain
length and to the average number of hydroxyl groups
esterified, but they do not provide information on the
distribution of products. To illustrate this, in figure 4 the
saponification numbers have been plotted versus their
corresponding hydroxyl numbers for a series of oleic
acid esters of polyglycerols. Figure 4 shows that many
different mixtures of esters can exhibit a given couple of
saponification and hydroxyl values, and that, especially
in the region of lower OH values, these mixtures could
contain a very large number of esters with different
esterification degree and polyglycerol chain length.
In table 2, we have collected some representative
examples on the manufacture of PGE, selected among

113

those reported both in journal and patent literature that


provide a more detailed composition of the ester
mixture.
Several procedures have been reported to increase
significantly the monoester yield. The transesterification
between polyglycerols and methyl esters of fatty acid has
been successfully achieved by using an emulsion medium
in the presence of alkaline catalysts (NaOH) [69].
Anionic emulsifiers, Na-stearate particularly, are very
effective in promoting the reaction between tetraglycerol
and methyl stearate. For a molar ratio of
polyol=acid 2, a selectivity to the monoester of 98%
at 100% conversion has been reported by using 10 wt%
(of the oil phase) of Na-stearate as emulsifier agent.
The reaction between glycidol and a fatty acid or
fatty monoglyceride has also been claimed to produce a
high yield of polyglycerol monoester [73]. The reaction is
carried out at temperatures around 150  C in the
presence of phosphoric acid as catalyst <0:5 wt%,
and the glycidol is slowly added to the mixture of fatty
acid and catalyst. Products containing more than 70% of
the monoester are obtained using different glycidol/fatty
acid ratios. However, there is no indication on the
polyglycerol chain-length distribution. Earlier reports
on this same reaction use Lewis acids, like BF3 and
SnCl4 , as catalysts [77].
To our knowledge, the use of porous, zeolite-like
materials as catalysts for the esterification of polyglycerols has not yet been reported. This is not surprising,
because of the low performance of these materials in the
selective esterification of the smaller glycerol molecule.
However, new opportunities for shape-selective catalysis
would be offered by the new generation of mesoporous
materials, as their functionalization with an appropriate

Table 2
Synthesis of fatty acid esters of polyglycerol
Polyglycerol esterification and transesterification
Fatty compound

Fatty
c./PGL
ratio

Capric acid

Lauric acid

1.1

Erucic acid
Methyl stearate
Methyl stearate

6
0.5
1

Catalyst
[wt%]

Reaction
temp.
(8C)

Fatty acid
conversion
(%)

Glycerol
units

Ester
composition

References

Dodecylbenzene sulfonic
acid H3PO2 [0.2]
Dodecylbenzene sulfonic
acid H3PO2 [0.2]
LiCl
NaOH [0.15]
NaOH [0.15]

145165

100

14 (88%)

50% mono-, 40% diester

[66]

145165

100

14 (88%)

35% mono-, 55% diester

[66]

230
130
130

100
89

6
4 (mean)
4 (mean)

75% mean esterifn.


98% monoester
56% monoester

[65]
[69]
[69]

Addition polymerization of glycidol to fatty compounds


Fatty compound

Lauric acid
Lauric acid

Fatty
c./glycidol
ratio

Catalyst
[wt%]

Reaction
temp.
(8C)

1/6
1/10

H3PO4 [0.02]
H3PO4 [0.02]

140
140

Fatty acid
conversion
(%)

Glycerol
units

Ester
composition

References

6 (mean)
10 (mean)

91% monoester
77% monoester

[73]
[73]

C. Marquez-Alvarez et al./Fatty acid esters of polyols

114

OH
O

OH

HO

OH

&
- H2O
HO
OH

OH

HO

OH

OH

2,5-sorbitan

1,4-sorbitan
OH

HO
OH

OH

sorbitol

-2 H2O

HO

OH
isosorbide
Scheme 5.

The sorbitol molecule (a hexitol) contains six hydroxyl groups that can be esterified with carboxylic acids.
However, this molecule dehydrates partially at moderate
temperature, forming as main products the hexitans
(mainly 1,4-sorbitan and minor amounts of 2,5-sorbitan), which contain four hydroxyl groups, and hexides
(mainly isosorbide), which contain two hydroxyl groups
only (scheme 5). Therefore, the result of the esterification process of sorbitol with fatty acyl groups is a
complex mixture of different compounds, whose nature
and relative proportions are governed by the specific
reaction conditions. As a consequence, the HLB value of
the esterification product is also strongly affected by the
conditions prevailing in the esterification process.
Sorbitan fatty acid esters can be produced by direct,
base- or acid-catalyzed reaction of sorbitol with fatty
acids at elevated temperature, or by transesterification
of sorbitol with triglycerides or fatty acid methyl esters
catalyzed by basic compounds [7887]. Homogeneous
catalysts, either acids ( p-toluenesulfonic acid, sulfuric
acid, phosphoric acid) or alkalis (NaOH, KOH, alkaline
carbonates) are generally used in the absence of
solvents. The use of a nitrogen atmosphere or vacuum
is required to remove the water and avoid excessive
oxidation of the reagents.
The anhydrization degree of sorbitol increases in the
presence of an acid, and therefore acid catalysts tend to
produce mixtures with a high proportion of esters of
sorbides, while alkaline catalysts lead mainly to esters of
sorbitans [78]. Unfortunately, most of the scientific
publications and patents only report the saponification
and hydroxyl values of the ester mixtures obtained. A
comparison of these values with the theoretical values
calculated for pure esters (as an example, in figure 5,

200

65
4
2

-1

4. Sorbitol monoesters

those of oleic acid esters as well as the nonesterified


polyols are reported) provides the average anhydrization
degree of the sorbitol and the average number of
hydroxyl groups esterified in these mixtures, but the
actual distribution of products cannot be obtained. In
table 3, we have collected some representative examples
of sorbitol esterification and transesterification processes for which the composition of the esters mixtures
obtained is reported.
A disadvantage of the base-catalyzed direct reaction
is that the product is usually highly colored [78].
Therefore, treatment with a bleaching agent such as
hydrogen peroxide and adsorbents is required to obtain
acceptable color values. Two-step processes have been
established where the acid-catalyzed anhydrization of
sorbitol takes place before the esterification or transesterification reaction [80,83,88,89]. It has also been
reported that lower color values are obtained when
reducing acids, like phosphorous and hypophosphorous
acids, are used as catalysts [83,86], or by addition of
reducing agents like sodium borohydride [83].

Saponification No. (mg KOHg )

acid or a base organic moieties would probably afford


active and selective catalysts, as it has been outlined
above for glycerol esterification.

3
2

3
2

es
ter
ific
a

150
1

tio
n

de
g

100

50

sorbitans

sorbides

sorbitol

500

ree

1000

1500

2000

-1

OH No. (mg KOHg )

Figure 5. Saponification and OH numbers of oleic acid esters of


sorbitol and its anhydrides. Figures indicate the esterification degree
(number of fatty acyl groups per molecule) of each ester.

C. Marquez-Alvarez et al./Fatty acid esters of polyols

115

Table 3
Synthesis of fatty acid esters of sorbitol and sorbitol anhydrides
Fatty compound

Fatty
c./sorbitol
ratio

Catalyst

Reaction
temp.
(8C)

Fatty acid
conversion
(%)

Ester
composition

References

Coconut oil fatty acids


Coconut oil fatty acids
Oleic acid
Oleic acid

1
1
1.45
1.47

235
225
200220
245

100
100

70% sorbitan monoester


Mostly sorbide monoester
Mostly sorbitan diestera
Mostly sorbitan diester

[78]
[78]
[83]
[86]

Lauric acid

NONE
H3PO4 (pH 2.0)
KOH
H3PO3 NaOH
(0.6 : 1 molar ratio) 2.6 wt%
Sulfonic acid functionalized
MCM-41

95% isosorbide diester

[85]

110

100
33

Sorbitol dehydration is carried out in a previous step using hypophosphorous acid as catalyst.

The reports on the use of heterogeneous solid


catalysts for sorbitol esterification are scarce. Van Rhijn
et al. [85] described the formation of dilauryl isosorbide
by reacting lauric acid and D-sorbitol (acid to polyol
molar ratio equal to 6) in the presence of a siliceous
MCM-41 catalyst functionalized with propylsulfonic
acid groups. At a lauric acid conversion of 33%, the
selectivity to the isosorbide diester was 95%. At lower
conversion (17%), isosorbide monolaurate is also
detected.
Interestingly,
the
large-pore
zeolite
Hbeta
Si=Al 12:5 is reported to be inactive under the same
reaction conditions [85], owing to the preferential
adsorption of the polyol on the zeolite. The better
performance of the SO3 H-MCM-41 material is attributed to its higher hydrophobicity. Therefore, it would be
expected that this catalyst, functionalized with hydrophobic alkyl groups, exhibits an improved activity in the
sorbitol esterification. However, the predominance of
isosorbide esters reveals that the acid catalyst favors the
deep sorbitol dehydration due to its strong acidity.
Furthermore, the hydrophilic character of the isosorbide
monoesters is low (very low HLB value), owing to their
low hydroxyl content, which makes the process unattractive for the synthesis of emulsifiers (nevertheless
isosorbide esters find application in other fields; for
example, the diester with 2-ethylhexanoic acid has been
claimed to be promising as a green PVC plasticizer [90]).
A hydrophobic catalyst containing acid centers of
medium strength would be more appropriate for the
direct esterification of sorbitol. Moreover, solid base
catalysts would also be useful as they will probably
decrease the anhydrization of the hexitol.
A more systematic exploration of the behavior of
different zeolites in the esterification of sorbitol with
oleic acid has been reported recently [91]. Zeolite Hbeta
Si=Al 13 turns to be active at 200  C (sorbitol/oleic
acid 1, 14:6 wt% of catalyst), but the ester selectivity
is similar to that of p-toluenesulfonic acid. Decreasing
the crystal size by one order of magnitude (to 12 nm)
does not change the activity and selectivity, but both

decrease when the Si/Al ratio is increased to 50. Zeolite


beta free of defects (i.e. free of silanol groups)
synthesized in fluoride medium is more active, which
has been attributed to its higher hydrophobicity. Zeolite
ITQ-2 (delaminated MCM-22) is also active in the
esterification reaction.
In all these examples, the OH content of the esters
was determined according to standard procedures, and a
value of 100 mg KOH g1 was obtained. Such a low OH
content of the product is consistent with the formation
of isosorbide esters and/or sorbitan esters having a high
degree of esterification (figure 5). Therefore, it can be
concluded that a low yield of monoesters is obtained
with these zeolitic catalysts for sorbitol conversion levels
in the range of 48 to 99%.
Cs-containing heteropolyacids (Hx Cs3x PW12 O40 ,
x 0:5 and 1) have also been used as esterification
catalysts [91]. At 150  C and a sorbitol/oleic acid molar
ratio of 1, with 13.6 wt% of catalyst, quantitative
conversion of both reagents is obtained. However, the
OH value of the product is about 110, which suggests
the predominance of isosorbide monoesters in the
mixture.
To summarize these results, it is concluded that no
sorbitol or sorbitan monoesters are obtained by using
these acid catalysts in the direct esterification of sorbitol
with oleic acid. In order to improve the selectivity to the
monoacyl derivatives, part of the hydroxyl groups of the
sorbitol were protected by forming the corresponding
ketals by reaction with acetone in the presence of HCl as
catalyst [91]. This modification increases the solubility
of sorbitol in oleic acid, which allows the esterification
reaction to take place at lower temperature 135  C.
Under these conditions, the oleic acid conversion level
obtained with the zeolite mordenite Si=Al 10 was
higher than 90% at 48 h, whereas the OH value of the
product was as high as 360. OH values of the esters
mixtures obtained with zeolites beta and ITQ-2 and
heteropolyacids also increased by means of the ketalization of sorbitol. Therefore, this procedure probably
reduces the transformation of sorbitol into isosorbide

116

C. Marquez-Alvarez et al./Fatty acid esters of polyols

and also decreases the average esterification degree of


the product.

with lower sorbitol anhydrization degree than those


obtained with acid catalysts.

5. Concluding remarks

Acknowledgments

The current technology for the production of fatty


acid esters of glycerol, polyglycerol and sorbitol suffers
from severe drawbacks due to the use of homogeneous
catalysts. The synthesis of these emulsifiers generates
large amounts of waste residues because the mineral
acids or alkalis that act as catalysts must be neutralized
and bleaching agents and adsorbents are usually needed
to obtain products with low color values. Owing to the
low selectivity of the homogeneous catalysts, a large
excess of polyol is needed in order to increase the yield
of the valuable esters, those with a low esterification
degree, therefore requiring separation of the reactant in
excess and increasing the energy demand. Furthermore,
costly molecular distillation of the glyceride mixtures
obtained must be carried out to obtain the required
concentration of monoglycerides. In the case of sorbitol
and polyglycerol, complex mixtures of esters with a wide
HLB distribution are being currently produced. However, ester mixtures with a more defined HLB would
exhibit improved emulsifying properties.
The potential of solid catalysts to overcome most of
these drawbacks has hardly been explored. A relatively
important research effort has been devoted to the study
of solid catalysts for the synthesis of monoglycerides.
Significant improvements have been reported by using
zeolitic catalysts, and, especially, mesostructured solids,
like siliceous MCM-41 structures, with acid functionalities. As the highest yields of monoglycerides that have
been obtained with these catalysts are still below those
required for commercial use of these esters mixtures,
there are still several issues that need further research.
The possibility to tune the pore size and generate
different pore architectures in these mesostructures can
allow the tailoring of the catalyst according to the fatty
acid used in order to optimize the monoester yield.
Different functionalities can be incorporated on the
surface of these mesostructures. Beneficial effects of the
control of the catalyst hydrophilic properties by attachment of alkyl chains on its surface have been reported
that deserve further study. Likewise, these mesostructures are interesting catalysts for the synthesis of esters
of sorbitol and polyglycerol. Acid-functionalized mesostructures are good candidates to be tested as catalysts
for the selective synthesis of polyglycerol monoesters at
low fatty acid to polyglycerol ratio. On the other hand,
base-functionalized mesostructures have been reported
to catalyze the etherification reaction of glycerol to
produce polyglycerol with a narrow chain-length distribution. These materials could be excellent catalysts
for the esterification or transesterification of sorbitol to
obtain esters of high HLB, as they would produce esters

The financial support of the Spanish CICyT (projects


MAT2000-1167-C02-02 and PETRI95-0537-OP), CAM
(project 07M/48/2002) and Betaqu mica, S.A. is gratefully acknowledged.

References
[1] W.C. Griffin, J. Soc. Cosmetic Chem. 1 (1949) 311.
[2] G. Dieckelmann and H.J. Heinz, The Basics of Industrial
Oleochemistry (Petr Pomp, Essen, 1988).
[3] U.T. Bornscheuer, Enzyme Microb. Technol. 17 (1995) 578.
[4] F.J. Plou, M. Barandiaran, M.V. Calvo, A. Ballesteros and E.
Pastor, Enzyme Microb. Technol. 18 (1996) 66.
[5] C. Otero, J.A. Arcos, M.A. Berrendero and C. Torres, J. Mol.
Catal. B 11 (2001) 883.
[6] W. Tsuzuki, Y. Kitamura, T. Suzuki and S. Kobayashi,
Biotechnol. Bioeng. 64 (1999) 267.
[7] D. Charlemagne and M.D. Legoy, J. Am. Oil Chem. Soc. 72
(1995) 61.
[8] A. Brock, B. Gruning and G. Hills, European Patent Appl. EP
1250917 (2003).
[9] N.O.V. Sonntag, J. Am. Oil Chem. Soc. 59 (1982) 795A.
[10] A. Corma, S. Iborra, S. Miquel and J. Primo, J. Catal. 173 (1998)
315.
[11] S. Bancquart, C. Vanhove, Y. Pouilloux and J. Barrault, Appl.
Catal. A 218 (2001) 1.
[12] A. Corma, H. Garc a, S. Iborra and J. Primo, J. Catal. 120 (1989)
78.
[13] S. Abro, Y. Pouilloux and J. Barrault, Stud. Surf. Sci. Catal. 108
(1997) 539.
[14] I. D az, A. Mart nez-Guerenu, F. Moh no, J. Perez-Pariente and
E. Sastre, Actas do XVII Simposio Ibero-americano de Catalise
(FEUP, Porto, 2000) p. 559.
[15] N. Sanchez, M. Mart nez and J. Aracil, Ind. Eng. Chem. Res. 36
(1997) 1524.
[16] E. Heykants, W.H. Verrelst, R.F. Parton and P.A. Jacobs, Stud.
Surf. Sci. Catal. 105 (1997) 1277.
[17] M. da Silva-Machado, J. Perez-Pariente, E. Sastre, D. Cardoso
and A. Mart nez-Guerenu, Appl. Catal. A 203 (2000) 321.
[18] M. da Silva-Machado, D. Cardoso and J. Perez-Pariente, Stud.
Surf. Sci. Catal. 130 (2000) 3417.
[19] Y. Pouilloux, S. Abro, C. Vanhove and J. Barrault, J. Mol. Catal.
A 149 (1999) 243.
[20] A. Corma, V. Fornes, M.T. Navarro and J. Perez-Pariente,
J. Catal. 148 (1994) 569.
[21] J. Perez-Pariente, I. D az, F. Moh no and E. Sastre, Appl. Catal.
A in press.
[22] W.D. Bossaert, D.E. de Vos, W.M. Van Rhijn, J. Bullen, P.J.
Grobet and P.A. Jacobs, J. Catal. 182 (1999) 156.
[23] I. D az, F. Moh no, J. Perez-Pariente and E. Sastre, Appl. Catal.
A 205 (2001) 19.
[24] I. D az, C. Marquez-Alvarez, F. Moh no, J. Perez-Pariente and
E. Sastre, Microporous Mesoporous Mater. 4445 (2001) 295.
[25] I. D az, C. Marquez-Alvarez, F. Moh no, J. Perez-Pariente and
E. Sastre, J. Catal. 193 (2000) 283.
[26] I. D az, C. Marquez-Alvarez, F. Moh no, J. Perez-Pariente and
E. Sastre, J. Catal. 193 (2000) 295.
[27] F. Moh no, I. D az, J. Perez-Pariente and E. Sastre, Stud. Surf.
Sci. Catal. 142 (2002) 1275.

C. Marquez-Alvarez et al./Fatty acid esters of polyols


[28] I. D az, F. Moh no, J. Perez-Pariente and E. Sastre, Appl. Catal.
A 242 (2003) 161.
[29] I. D az, F. Moh no, M.T. Blasco, E. Sastre and J. Perez-Pariente, 4th
International Mesostructured Materials Symposium, submitted.
[30] K. Wilson, A.F. Lee, D.J. Macquarrie and J.H. Clark, Appl.
Catal. A 228 (2002) 151.
[31] M. Boveri, J. Agundez, I. D az, J. Perez-Pariente and E. Sastre,
Actas de la Reunion de la SECAT (Malaga-Torremolinos, 2003).
[32] I. D az, F. Moh no, E. Sastre and J. Perez-Pariente, Stud. Surf.
Sci. Catal. 135 (2001) 1383.
[33] Y. Sakamoto, I. D az, O. Terasaki, D. Zhao, J. Perez-Pariente,
J.M. Kim and G.D. Stucky, J. Phys. Chem. B 106 (2002) 3118.
[34] I. D az, F. Moh no, J. Perez-Pariente, E. Sastre, P.A. Wright and
W. Zhou, Stud. Surf. Sci. Catal. 135 (2001) 1248.
[35] A. Cauvel, G. Renard and D. Brunel, J. Org. Chem. 62 (1997)
749.
[36] X. Lin, G.K. Chuah and S. Jaenicke, J. Mol. Catal. A 150 (1999)
287.
[37] N. Garti, A. Aserin and B. Zaidman, J. Am. Oil Chem. Soc. 58
(1981) 878.
[38] E. Symmes, U.S. Patent Appl. U.S. 1,696,337 (1928).
[39] H.A. Bruson, U.S. Patent Appl. U.S. 2,284,127 (1942).
[40] K. Babayan, U.S. Patent Appl. U.S. 3,637,774 (1972).
[41] R.T. McIntyre, J. Am. Oil Chem. Soc. 56 (1979) 835A.
[42] J. Van Heteren, P. Cornelis, F. Reckweg and M.F. Stewart,
European Patent Appl. EP 0070080 (1983).
[43] J. Klein, P. Daute, U. Hees and B. Beuer, PCT Int. Appl. WO
9302124 (1993).
[44] G. Hillion and R. Stern, European Patent Appl. EP 0518765
(1992).
[45] European Parliament and Council Directive No. 95/2/EC.
[46] Code of Federal Regulation, Vol. 3, Title 21, Food and Drug
Administration, US Government (2002).
[47] H. Helfert, A. Hettche,S. Weiss, W.D. Balzer, W. Dietsche and G.
Hatzmann, German Patent Appl. DE 3118417 (1982).
[48] K.S. Dobson, K.D. Williams and C.J. Boriack, J. Am. Oil Chem.
Soc. 70 (1993) 1089.
[49] S. Akimoto, Japanese Patent Appl. JP58198429 (1983).
[50] S. Uda and E. Takemoto, Japanese Patent Appl. JP61140534
(1986).
[51] S. Uda and E. Takemoto, Japanese Patent Appl. JP61043627
(1986).
[52] U. Hees, R. Bunte, J.W. Hachgenei, P. Kuhm and E.G. Harris,
PCT Int. Appl. WO9325511 (1993).
[53] K. Cottin, J.M. Clacens, Y. Pouilloux and J. Barrault, Ol. Corps
Gras, Lipides 5 (1998) 407.
[54] J.M. Clacens, Y. Pouilloux, J. Barrault, C. Linares and M.
Goldwasser, Stud. Surf. Sci. Catal. 118 (1998) 895
[55] J.M. Clacens, Y. Pouilloux and J. Barrault, Appl. Catal. A 227
(2002) 181.
[56] T. Nakamura and M. Yamashita, U.S. Patent Appl. U.S. 2002/
0035238 (2002).
[57] N. Garti, G.F. Remon and B. Zaidman, Tenside Deterg. 23 (1986)
6.
[58] S.B. Harvey, U.S. Patent Appl. U.S. 5,585,506 (1996).

[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]

[90]
[91]

117

D. Lemke, PCT Int. Appl. WO 0236534 (2002).


UK Patent Appl. GB1025265 (1966).
M. Ishitobi, U.S. Patent Appl. U.S. 2002/123439 (2002).
H. Tomokazu and Y. Akifumi, Japanese Patent Appl. JP 8217725
(1996).
P. Seiden and R.A. Woo, Patent Appl. CA 1185610 (1985).
B.K. Tjurin, German Patent Appl. DE 2511807 (1976).
K. Matsuda and M. Kitao, U.S. Patent Appl. U.S. 5,773,073
(1998).
G. Jakobson, W. Siemanowski and K.H. Uhlig, U.S. Patent Appl.
U.S. 5,424,469 (1995).
P. Denecke, G. Borner and V. Von Allmen, UK Patent Appl. GB
2073232 (1981).
G. Jakobson, W. Siemanowski and K.H. Uhlig, U.S. Patent Appl.
U.S. 5,466,719 (1995).
V.R. Kaufman and N. Garti, J. Am. Oil Chem. Soc. 59 (1982)
471.
I. Beseda and P.E. de Detrich, U.S. Patent Appl. U.S. 4,950,441
(1990).
V. Eychenne, L. Debrauwer and Z. Mouloungui, J. Surfact.
Deterg. 3 (2000) 173.
E. Ulsperger, L. Richter and F. Mainas, Patent Appl. DD 41661
(1965).
T. Endo and T. Daito, European Patent Appl. EP0758641 (1997).
T. Endo and K. Maruo, Japanese Patent Appl. JP 2000143794
(2000).
T. Endo and T. Ueno, Japanese Patent Appl. JP 2000143584
(2000).
K. Maruo and H. Omori, Japanese Patent Appl. JP 11222521
(1999).
G. Vanlerberghe and H. Sebag, U.S. Patent Appl. U.S. 4,515,775
(1985).
K.R. Brown, U.S. Patent Appl. U.S. 2,322,820 (1943).
L.D. DHinterland, G. Mouzin and H. Cousse, French Patent
Appl. FR 2176487 (1973).
G.J. Stockburger, U.S. Patent Appl. U.S. 4,297,290 (1981).
T. Masuda, M. Honjiyou and K. Watanabe, Japanese Patent
Appl. JP 57133197 (1982).
S. Ropuszynski and E. Sczesna, Tenside Deterg. 22 (1985) 190.
N. Milstein, PCT Int. Appl. WO 9200947 (1992).
G.O. Gruetzmacher, J.W. Raggon and B. Wlodecki, U.S. Patent
Appl. U.S. 5,612,080 (1997).
W.M. Van Rhijn, D.E. De Vos, B.F. Sels, W.D. Bossaert and
P.A. Jacobs, Chem. Commun. (1998) 317.
J.M.H. Ellis, J.J. Lewis and R.J. Beattie, PCT Int. Appl. WO
9804540 (1998).
PCT Int. Appl. WO 0128961 (2001).
L. Szotyori, A. Ujhidy, I. Szabo, G. Bozoki and J. Bathory,
Hungarian Patent Appl. HU 5159 (1972).
R.V. Vladea, M.A. Vineze, T.L. Simandan, L.M. Rusnac, A.T.
Fenesiu and M. Mosincat, Romanian Patent Appl. RO 87381
(1985).
H. Luitjes and J.C. Jansen, PCT Int. Appl. WO 9945060 (1999).
A. Velty, Ph.D. Thesis (Universidad Politecnica de Valencia,
2003).

También podría gustarte