ArticlePDF AvailableLiterature Review

Sucrose fatty acid esters: synthesis, emulsifying capacities, biological activities and structure-property profiles

Taylor & Francis
Critical Reviews In Food Science and Nutrition
Authors:

Abstract and Figures

The notable physical and chemical properties of sucrose fatty acid esters have prompted their use in the chemical industry, especially as surfactants, since 1939. Recently, their now well-recognized value as nutraceuticals and as additives in cosmetics has significantly increased demand for ready access to them. As such a review of current methods for the preparation of sucrose fatty acid esters by both chemical and enzymatic means is warranted and is presented here together with an account of the historical development of these compounds as surfactants (emulsifiers). The somewhat belated recognition of the antimicrobial, anticancer and insecticidal activities of sucrose esters is also discussed along with a commentary on their structure-property profiles.
Content may be subject to copyright.
Full Terms & Conditions of access and use can be found at
https://www.tandfonline.com/action/journalInformation?journalCode=bfsn20
Critical Reviews in Food Science and Nutrition
ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/bfsn20
Sucrose fatty acid esters: synthesis, emulsifying
capacities, biological activities and structure-
property profiles
Yinglai Teng , Scott G. Stewart , Yao-Wen Hai , Xuan Li , Martin G. Banwell &
Ping Lan
To cite this article: Yinglai Teng , Scott G. Stewart , Yao-Wen Hai , Xuan Li , Martin G. Banwell
& Ping Lan (2020): Sucrose fatty acid esters: synthesis, emulsifying capacities, biological
activities and structure-property profiles, Critical Reviews in Food Science and Nutrition, DOI:
10.1080/10408398.2020.1798346
To link to this article: https://doi.org/10.1080/10408398.2020.1798346
Published online: 04 Aug 2020.
Submit your article to this journal
View related articles
View Crossmark data
REVIEW
Sucrose fatty acid esters: synthesis, emulsifying capacities, biological activities
and structure-property profiles
Yinglai Teng
a,b
, Scott G. Stewart
c,d
, Yao-Wen Hai
a
, Xuan Li
a
, Martin G. Banwell
a,d,e
, and Ping
Lan
a,b,d
a
Institute for Advanced and Applied Chemical Synthesis, Jinan University, Zhuhai, Guangdong, China;
b
College of Pharmacy, Jinan University,
Guangzhou, Guangdong, China;
c
School of Molecular Sciences, The University of Western Australia (M310), Crawley, Western Australia,
Australia;
d
Research Laboratories, Guangzhou Cardlo Biochemical Technology Co., Ltd, Guangzhou, Guangdong, China;
e
Research School of
Chemistry, Institute of Advanced Studies, The Australian National University, Canberra, Australian Capital Territory, Australia
ABSTRACT
The notable physical and chemical properties of sucrose fatty acid esters have prompted their use
in the chemical industry, especially as surfactants, since 1939. Recently, their now well-recognized
value as nutraceuticals and as additives in cosmetics has significantly increased demand for ready
access to them. As such a review of current methods for the preparation of sucrose fatty acid
esters by both chemical and enzymatic means is warranted and is presented here together with
an account of the historical development of these compounds as surfactants (emulsifiers). The
somewhat belated recognition of the antimicrobial, anticancer and insecticidal activities of sucrose
esters is also discussed along with a commentary on their structure-property profiles.
KEYWORDS
Biological activities;
chemical and enzymatic
synthesis; emulsifying
properties; esterification
and transesterification;
structure-property profiles
Introduction
The disaccharide sucrose (b-D-fructofuranosyl a-D-glucopyr-
anoside), commonly called table sugar, is obtained from
various plant sources, is easily refined and is marketed
world-wide. Global production of this compound reached
ca. 180 million metric tons in 2016/2017 (Canadian Sugar
Institute 2019a). Direct consumption (as a food) accounted
for most of the demand but sucrose is also employed as a
starting material in the production of emulsifying agents,
detergents, artificial sweeteners and various other derivatives
(Human Metabolome Database 2019). Sucrose, which is
photosynthesized by almost every green plant but is most
readily derived from sugarcane and sugar beet (Canadian
Sugar Institute 2019b), is also exploited in the production of
ethanol and certain fine chemicals (Deneyer, Ennaert, and
Sels 2018) including sucrose fatty acid esters (Neta, Teixeira,
and Rodrigues 2015).
Sucrose fatty acid esters, as with esters more generally,
are formally derived from the constituent alcohol (viz.
sucrose) and (naturally occurring) fatty acid. Notable
commercially available forms of such esters include the
mono-esters derived from, inter alia, steric acid, palmitic
acid, lauric acid, oleic acid, erucic acid and myristic acid.
These esters, which frequently come as admixtures with
various of their di-, tri- and tetra-acylated counterparts
(Mitsubishi Chemicals 2019; Sisterna 2019a; Sisterna 2019b),
have excellent emulsifying properties as well as a range of
other useful physical properties that follow from their
amphipathic nature (Szuts et al. 2010; Farr
an et al. 2015;
Soultani et al. 2003). Importantly, such nonionic surfactants
are considered green because they are both nontoxic and
biodegradable. Furthermore, both the constituent sucrose
and fatty acid residues are inexpensive and renewable agri-
cultural products (Szuts et al. 2010; Farr
an et al. 2015;
Soultani et al. 2003). These various attributes mean that
sucrose esters have become profoundly important commod-
ity chemicals in the food (common ingredient number E473,
according to Food Standards Agency 2018), nutraceutical,
cosmetic, dental and pharmaceutical industries (Szuts et al.
2010; Farr
an et al. 2015; Soultani et al. 2003). Even a decade
ago, the world production of sucrose esters was estimated to
be above 6000 t/year and these were selling for 4 to 20
USD/kg. While mainly employed as food additives, they also
feature in the manufacture of cosmetics, pharmaceuticals
and detergents (Otomo et al. 2009). In 2015 the global
sucrose ester market was valued at USD 55.7 million and
projected to reach USD 74.6 million in 2020
(MarketsandMarkets 2019b). Key manufacturers of these
now much sought-after community chemicals include
Mitsubishi-Kagaku, Croda, Dai-ichi-Kogyo Seiyaku,
Goldschmidt, Sisterna, Weixi Spark, Evonik Industries and
BASF (Hill and Rhode 1999; Cruces et al. 2001;
MarketsandMarkets 2019a; Parker, Khan, and York 1973).
The methods of synthesis of sucrose-based esters was the
subject of a concise review in 2001 (Polat and Linhardt
2001) and surveys of enzymatic approaches have followed
(Shi, Li, and Chu 2011; Gumel et al. 2011). In the present
article we delineate both industrial and laboratory-based
CONTACT Ping Lan ping.lan@jnu.edu.cn Institute for Advanced and Applied Chemical Synthesis, Jinan University, Zhuhai, Guangdong 510632, China.
ß2020 Taylor & Francis Group, LLC
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION
https://doi.org/10.1080/10408398.2020.1798346
processes for the formation of the title esters that have
emerged in the intervening period. In addition, a commen-
tary on the emulsifying properties as well as the antimicro-
bial, antitumor and insecticidal activities of these important
esters is provided. Structure-property profiles arising from
an examination of these follows.
Chemical synthesis
Many simple esters are prepared by esterification under
Fischer-type conditions from the relevant acid and alcohol
but sucrose esters are not readily formed by such direct
means (Parker, Khan, and Mufti 1976). This is because
sucrose undergoes glycosidic bond cleavage under the neces-
sary conditions (viz.atpH23 and 7080 C) and so pro-
ducing a mixture of glucose and fructose (which is called
inverted or invert sugar) (Soares et al. 2019). Accordingly,
the industrial production of sucrose esters normally involves
the transesterification of fatty acid esters with sucrose and
so circumventing the need to employ strongly acidic condi-
tions (Parker, Khan, and Mufti 1976).
Establishing a single-step, inexpensive and regioselective
chemical synthesis of sucrose fatty acid esters is particularly
challenging given there are eight distinct hydroxyl groups
associated the parent carbohydrate (Queneau et al. 2008).
The per-ester sucrose octaacetate was prepared as early as
1880 by Herzfeld (1880) and in 1939 Cantor (1939) patented
a method for forming sucrose fatty acid esters from starch
factory by-products and simultaneously claimed that such
products could be used as emulsifying agents or fats. Of
course, there are many possible esters (255 in fact) available
by combining, through esterification, any given acid with
sucrose, including 8 mono-esters, 28 di-esters, 56 tri-esters
and so on (Weiss et al. 1971). Given the mono-esters of
sucrose are the most desirable as emulsifiers because of their
generally more appropriate HLB values (see below) (Polat
and Linhardt 2001), methods for the selective production of
these have been the focus of considerable attention. Of the
eight distinct hydroxyl groups embodied within sucrose it
has been established, unsurprisingly perhaps, that acylation
occurs preferentially at the primary ones and with some
selectivity being observed between these, viz. at the 6-OH
(glucose unit), then at the 60-OH (fructose unit) and, finally,
at the 10-OH (fructose unit) positions (Haines 1976).
However, the difference in reactivity of these centers is not
such that mono-esterification is readily achieved and a
detailed evaluation of the composition of sucrose esters
from various commercial sources (Table 1) emphasizes this
point (Cruces et al. 2001). In this study a large variation in
mono- and di-ester ratios was observed, a feature that
reflects the differing synthetic protocols used to make these
compounds as well as the potential for equilibration between
certain of them. Related studies are notable for their use of
quantitative GC and HPLC techniques to achieve the separ-
ation of the component esters and the often effective detec-
tion of these using evaporative light scattering (ELS) and/or
refractive index techniques (Moh, Tang, and Tan 2000).
Current industrial syntheses
Methods for the large (industrial) scale production of
sucrose esters traditionally involves the transesterification of
sucrose using a suitable triglyceride, a process normally con-
ducted in the presence of a catalyst such as potassium car-
bonate or a potassium soap. N,N-Dimethyformamide (DMF)
is the often used solvent (Parker, Khan, and York 1973) but
given its cost, toxicity and other safety issues, reactions
employing DMSO have also been explored. Under favorable
circumstances, mixtures comprising greater than 50%
sucrose mono-ester can be obtained with the remaining
components being other esters and the starting materials
(Figure 1) (Polat and Linhardt 2001). Challenges remain,
however, since removal of these high-boiling solvents
(153 C for DMF and 189 C for DMSO) (Royal Society of
Chemistry 2019) is a non-trivial matter and compounded by
the relatively low thermal stability of the product esters. The
necessary separation of by-products, so as to raise the
mono-ester component to levels acceptable to the market
(Guti
errez et al. 2018), is problematic and also creates waste
streams and so increasing production costs. Sucrose mono-
esters have also been produced industrially from fatty acid
methyl esters using reversible transesterification process with
attendant removal of co-produced methanol so as to drive
the reaction forward. Under the optimal conditions this
protocol can provide material containing upwards of 70% of
the mono-ester (Polat and Linhardt 2001).
Solvent-free processes have also been developed and used
in industry. So, for example, in 1970, Feuge et al. (1970)
found that the so-called interesterification of a suitable fatty
acid methyl ester with molten sucrose in the presence of
catalytic quantities of soap (usually a potassium soap) at
170185 C can produce the required mono-esters. However,
this approach has drawbacks since at the required (elevated)
temperatures sucrose only remains a liquid for brief periods
(several minutes) and when reduced pressures are applied to
remove the co-formed methanol then degradation of the
sugar occurs as evidenced by the significant darkening of
the reaction mixture. This impacts on both throughput and
product quality while the need to remove the catalyst adds
further technical difficulties (and attendant financial bur-
dens). Another solvent-free and large-scale process was
developed by Parker, Khan, and York (1973) and wherein
Table 1. Composition of commercial sucrose esters as determined by Plou
and coworkers (Cruces et al. 2001).
Compound Source
Composition (by wt.%)
Sucrose Mono-esters Di-esters
Sucrose caprylate Calbiochem 2 95 3
Sucrose caprate Sigma 5 93 2
Sucrose laurate Fluka 2 95 3
Sucrose laurate L70-C Sisterna 11 50 39
Sucrose laurate L-1695 Mitsubishi-Kagaku 2 77 21
Sucrose laurate L-595 Mitsubishi-Kagaku 0 27 43
Sucrose myristate M-1695 Mitsubishi-Kagaku 1 77 22
Sucrose palmitate P-1670 Mitsubishi-Kagaku 0 64 22
Sucrose palmitate P-1570 Mitsubishi-Kagaku 0 59 29
Sucrose palmitate P-170 Mitsubishi-Kagaku 2 88 10
Sucrose palmitate Sigma 9 65 26
Sucrose stearate S-1670 Mitsubishi-Kagaku 0 54 20
2 Y. TENG ET AL.
fine particles of sucrose are reacted with triglycerides (in the
form of oils such as tallow, palm oil or coconut oil) in the
presence of catalytic amounts of potassium carbonate. In
this way, the reaction can be performed at atmospheric pres-
sure and temperatures of around 125 C. Furthermore, the
reaction can be accelerated by adding emulsifiers (e.g. a
diglyceride). However, under such conditions a mixture of
sucrose esters and glycerides is formed. While attractive in
principle, the challenges (as mentioned above) attending the
solvent-free production of sucrose mono-esters mean that
DMF continues to be employed, as a matter of course, in
many commercial processes.
Laboratory-scale syntheses
Nature and selectivity of acylating agents
Approaches to sucrose esters that avoid traditional and nor-
mally ineffective (see above) Fischer esterification
approaches (Figure 2a) abound but it is clear that pure
mono-ester formation cannot be achieved by simply control-
ling the fatty acid ester stoichiometry in transesterification
reactions (Figure 2b). Furthermore, acyl group migration
can occur during the course of such reactions and/or during
isolation and thus introducing additional components into
the product mixture (Queneau, Fitremann, and Trombotto
2004). Sucrose ester syntheses using acyl chlorides (Figure
2c) have been studied with, for example, the esterification of
sucrose by octanoyl chloride at pH 10 shown to provide
mixtures of mono-octanoates, di-octanoates and tri-octa-
noates. By employing 4-(N,N-dimethylamino)pyridine
(DMAP) as an acylation catalyst under dilute conditions,
and by exploiting hydrophobic effects, greatly enhanced pro-
portions of mono-octanoates can be obtained (Th
evenet
et al. 1997). However, the need to strictly control the water
content of the reaction medium while using such protocols
as well as the high cost of catalysts such as DMAP mean
they are likely to be less attractive in industrial settings
(Kea, Walker, and Kline 1987;Th
evenet et al. 1999).
A range of replacements for acid chlorides as the acylat-
ing agent, including acid anhydrides and enol esters, has
been explored. So, for example, the Plusquellec Group
effected regioselective acylation at the 60-OH group of
sucrose using various 3-acyl-5-methyl-1,3,4-thiadiazole-
2(3H)-thiones (Figure 2d) in conjunction with diazobicyclo-
[2.2.2]octane (DABCO) (as base) at low temperature and
DMF as solvent (Chauvin and Plusquellec 1991).
Intriguingly, the same group found that by using 3-acyl-thia-
zoledine-2-thiones in the presence of NaH then a higher
yielding reaction took place and now at the C2-OH group
of sucrose (Chauvin, Baczko, and Plusquellec 1993). The
Queneau Group completed a detailed study on the distribu-
tion of sucrose mono-acylation products obtained on using
N-decanoylthiazolidinethione (Figure 2d) and various homo-
logues as electrophiles (Molinier et al. 2003). Dipotassium
hydrogenophosphate proved to be an effective base in these
cases and it was concluded that N-acylthiazolidinethiones
are better acylating agents than cyanoethyl-, methyl-
thioethyl- and methyl-esters. In another study, 6-O-acylated
sucrose derivatives were accessed in a highly selective man-
ner through reaction (in DMF) of the parent sugar-derived
dibutylstannylene acetal dimers with fatty acid anhydrides
(Vlahov, Vlahova, and Linhardt 1997; Bazin, Polat, and
Linhardt 1998). Subsequently the same sorts of intermediates
were used to regioselectively produce 6-O-acylsucroses and
6,30-di-O-acylsucroses (Wang, Zhang, and Yang 2007) but
such attractive outcomes are offset by the possibility of
product contamination by toxic tin residues. The three-fold
benzoylation of sucrose using benzoyl chloride in the pres-
ence of pyridine has been shown to occur selectively at the
6-O,1
0-Oand 6-Opositions although various other tri-ben-
zoates and certain tetra-benzoates are also produced (Clode
et al. 1985).
Irreversible transesterification reactions involving vinyl
esters (Figure 2d) provide a unique approach to the forma-
tion of sucrose mono-esters with vinyl acetate having been
used to acetylate sucrose. This reaction was reported to take
place in water under basic conditions at slightly elevated
temperatures although the regioselectivity of the process is
not clear (Smith and Tuschhoff 1957). In 2001, Plou et al.
reported the transesterification of sucrose using vinyl capryl-
ate, laurate, myristate or palmitate under strictly anhydrous
conditions in DMSO with Na
2
HPO
4
serving as the catalyst
(Cruces et al. 2001). By such means, remarkably high yields
Figure 1. The most common methods used for the industrial preparation of sucrose mono-esters (Parker, Khan, and York 1973; Polat and Linhardt 2001).
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 3
(8697%) of product mixtures rich in mono-esters
(9094 wt.% as determined by HPLC) were realized, an out-
come that compared favorably with industrial processes in
operation at the time. The predominant product of such
reactions was the 2-O-acylated material but the co-produc-
tion of its 3-O-acyl counterpart, arising through acyl group
migration, could not be avoided.
Much research into solvent-freetransesterification reac-
tions has focused on the solubilityof sucrose in the prod-
uct ester over the temperature range of 40140 C with
substrate decomposition only becoming evident above this
upper limit. Notably, the solubility of sucrose in the methyl
esters of various fatty acids was observed to increase from
0.25 to 1.40 g/100 g following addition of 2.5 wt.% sodium
stearate and even more so in presence of the surfactant
S1570 (Zhao et al. 2014).
Selective di-esterification at the 6-Oand 60-Opositions of
sucrose can be achieved using gallic acid as the nucleophile
under Mitsunobu conditions (Potier et al. 1999). In these
cases it is assumed that the surcrose-derived oxyphospho-
nium ion, prepared using PPh
3
and DIAD, undergoes
nucleophilic displacement (and the product esters were
examined for their capacities to serve as antioxidants).
Other Mitsunobu-based approaches have provided 6-O-pal-
mitoylsucrose (47% yield) (Bottle and Jenkins 1984) and the
corresponding 6-O-polyfluoroalkanoates (Abouhilale,
Greiner, and Riess 1991). 6-O-Acyl-30,60-anhydro- and 6-O-
acyl-30,40-anhydro-sucroses have been formed as by-products
during the esterification of unprotected sucrose by related
means (Molinier et al. 2004). These epoxide-containing
compounds can also be formed deliberately by using
Mitsunobu protocols (Guthrie, Jenkins, and Yamasaki 1980).
The esterification of sucrose and certain of its congeners
through coupling with activated sulfonic acid derivatives has
also been the subject of several studies and variations in the
reaction conditions provide complementary outcomes in
terms of regioselectivity (Bazin, Polat, and Linhardt 1998;
Kowalski, Cmoch, and Jarosz 2014; Wuts and Greene 2006;
Buchanen and Cummerson 1972; Teranishi 2002). Thus,
sucrose 2-O-p-toluenesulfonate can be obtained under one
set of conditions, while modest changes to these provides
access to the 6-Oor 60-Oregioisomers (Teranishi 2002).
The effect of bases
As suggested by the outcomes of certain of the reactions
described above, the choice of base employed in sucrose
esterification reactions can have a significant influence on
the yield of such processes. While both NaOH and KOH
can be useful for such purposes, in water and other protic
solvents their use leads to saponification of the desired ester.
As such they are only effective under strictly anhydrous con-
ditions, a requirement that limits their utility in industrial
settings. The milder potassium carbonate is a more suitable
catalytic base with a recent study of its application in the
esterification of sucrose using coconut esters in DMF focus-
ing on quantitating the most effective molar ratio of sucrose
to fatty acids, catalyst concentration, solvent volume, reac-
tion temperature and viscosity (Deshpande et al. 2013).
Sonication has been explored as a means for promoting the
esterification of sucrose and under appropriate conditions
good yields of mixtures of sucrose mono-esters can be
obtained. For example, the (trans)esterification of sucrose in
DMSO at 70 C using ethyl palmitate and K
2
CO
3
was
achieved in 73% yield to give a product mixture containing
92% of a ca. 2.5:1 mixture of the 6-O- and 60-O-acylated
mono-esters (Huang et al. 2010a). Related transesterifica-
tions employing K
2
CO
3
in methanol have been reported
(Vassilev et al. 2016).
Combinations of the acetylating agent N-decanoylthiazoli-
dinethione and various carbonates such as K
2
CO
3
,Li
2
CO
3
and Cs
2
CO
3
or phosphates (e.g.,K
2
HPO
4
) have been trialed
Figure 2. Procedures used for the production of sucrose esters: (a) esterification; (b) transesterification; (c) acyl chloride-based esterification and d) examples of
traditional and modern acylating agents (Chauvin and Plusquellec 1991; Chauvin, Baczko, and Plusquellec 1993; Molinier et al. 2003).
4 Y. TENG ET AL.
at ambient temperatures for effecting the esterification of
sucrose. A 71% yield of the 2-O-acylated mono-ester was
achieved using two equivalents of K
2
HPO
4
while various
mixtures of the 3-O,3
0-O4-O(diester), 2-O,4
0-O,6-Oand
10-O,6
0-Ocongeners were also observed with the precise
distribution being dependant on the base used (Molinier
et al. 2003). Some preference for formation of the 2-O,3-O-,
and 6-O-alkanoyl sucroses was seen. So when, for example,
lithium carbonate was employed selective esterification at
the 2-OH group was observed.
Acyl migration methods and sucrose ester stability
Both inter- and intra-molecular acyl migration protocols,
viz. transesterifications, play a pivotal role in the derivatiza-
tion of carbohydrates including sucrose. NMR techniques
involving
13
C-labeled substrates have been used to study
acyl group migrations within 1b-O-acyl glucuronides (Figure
3) (Akira et al. 1998) and, as might be expected, such migra-
tions most likely involve intramolecular nucleophilic attack
by the adjacent hydroxyl group at the carbonyl moiety and,
thereby, formation of a tetrahedral intermediate that itself
collapses to form the migrated product. Depending upon its
strength, added base can facilitate migration through accel-
eration of one or other of the discrete steps shown in Figure
3and under suitable conditions such acyl migrations can
provide a useful means for obtaining single regioisomeric
forms of carbohydrate mono-esters. Since 6-O-acylated
sucroses and their 10-Oand 60-Ocounterparts are the most
stable of the eight possible mono-esters (Molinier et al.
2003), these can be prepared by treating, for example, the
isomeric 2-O-acylsucroses with a (hindered) organic base
and so effecting O-acyl migration in up to 60% yield. Such
processes have been exploited to prepare both 6-O-octanoyl-
sucrose and 6-O-stearoylsucrose (Baczko et al. 1995). In an
extensive study of the isomerization of 3-O-monolaurate
sucrose acyl under basic conditions, the rate of migration
from the 3- to the 6-position was found to accelerate in the
presence of water both during the aqueous workup and
under the HPLC conditions used for analysis (Molinier et al.
2003). Notably, no significant isomerization of the substrate
occurred when this was treated with 1,8-diazabicy-
clo[5.4.0]undec-7-ene (DBU) in the absence of water.
While acyl migrations from the 2-Oto the 3-Opositions
are reversible, at equilibrium the 3-O-ester is normally
favored (Cruces et al. 2001; Baczko et al. 1995). Migration
from the 3-Oto the 6-Oposition can be observed and a
detailed study involving the use of various acylating agents,
including acid chlorides and alkyl chloroformates, estab-
lished that a range of factors, including steric ones, impact
on mono-ester product distributions (Th
evenet et al. 1999).
If appropriate control of various physical and chemical
parameters is attained then pure mono-esters can often
be generated.
Summary
While various laboratory-scale protocols have been estab-
lished for the reliable and selective production of sucrose
mono-esters, the translation of these into industrial settings
is challenging, not least because solvents such as pyridine
cannot be used in the food industry while acylating agents
such as acyl chlorides and vinyl esters are prohibitively
expensive and/or are environmentally disadvantageous. Such
issues have attracted the recent attention of various aca-
demic groups who acknowledge industrial imperatives and
have thus pursued the use of enzymes to prepare sucrose
esters. The nature and outcomes of such studies are
now discussed.
Enzymatic syntheses
Enzyme-catalyzed syntheses of sucrose esters offer a pathway
to higher purity forms of these products as a result of their
enhanced regioselectivities. Furthermore, these biocatalytic
processes are generally environmentally-friendly ones
because of the mild conditions involved and the nontoxic
nature of the (often aqueous) media involved (Franssen
et al. 2013). To date, however, no commercial sucrose ester
production process appears to exploit this enzyme-based
approach, not least because of the associated costs and the
need for these to be run in batch rather than flow mode
(Satyawali, Vanbroekhoven, and Dejonghe 2017). Both
lipases and proteases can be used for the esterification of
sucrose (Chang and Shaw 2009). The associated processes
have also been employed extensively for the production of
Figure 3. Acyl migration processes observed within the glucuronide framework (Akira et al. 1998).
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 5
other sugar esters, including lactose monolaurate (Walsh
et al. 2009). Detailed studies of such transformations, some
outcomes of which are discussed in the following sections,
reveal that the nature of the solvent, the water content of
the medium and the temperature as well as the method of
mixing are often crucial. Such requirements are dictated by
the need to retain/preserve/protect the enzyme active site
(viz. to avoid denaturing of the enzyme).
Lipases
Lipases are a group of esterases that catalyze the hydrolysis
of lipids and such enzymes are comprised of several sub-
classes including phospholipases and acylglycerol lipases
(Schomburg et al. 2017). In the present context, the most
important lipases are, arguably, the triacylglycerol lipases
(EC 3.1.1.3) that selectively hydrolyze insoluble triacylglycer-
ols at the substrate-water interface (BRENDA 2019b). As
such these lipases, especially immobilized forms, have found
a multitude of applications in industry particularly for flavor
and biodiesel production as well as in the treatment of waste
water (Filho, Silva, and Guidini 2019). Various lipases have
also been used for the laboratory-scale preparation of
sucrose esters, including certain commercially available and
immobilized ones such as Lipozyme TL IM (a lipase derived
from the thermophlic fungus Thermomyces lanuginosus, for-
mally Humicola lanuginosa) (Ferrer et al. 2005) and
Novozym 435 (derived from the yeast Candida antarctica B)
(Shao et al. 2018; Ye, Hayes, and Burton 2014).
Generally, when a lipase is hydrated then only small
amounts of water need be present in the reaction medium
so as to ensure its catalytic viability (Goderis et al. 1987).
On the other hand, since lipases function in equilibrium
(Figure 4a), hydrolysis of the product ester can occur even
when anhydrous (organic) solvents are employed (Gumel
et al. 2011). Thus, determining the optimal water content of
the reaction medium is a major consideration in any associ-
ated methodological studies (Chamouleau et al. 2001; Gumel
et al. 2011). Unlike many enzymes that follow
MichaelisMenten kinetics, lipase acyl-transfer reactions are
thought to proceed through a bi-bi ping-pong mechanism
(Figure 4b) with competitive substrate inhibition being
caused by the alcohols used as acyl acceptors (Martinelle
and Hult 1995; Reyes-Duarte et al. 2005).
Solvent effects
Lipases remain viable at low concentrations in a range of
organic solvents including n-hexane, DMSO, methanol,
xylene and cyclohexane (Salihu and Alam 2015). So, for
example, Novozym 435 has been deployed in the prepar-
ation of long-chain long fatty acid esters by treating sucrose
with the relevant acylating agent in a mixture of DMSO and
2-methyl-2-butanol as so to produce sucrose palmitate
(Reyes-Duarte et al. 2005). While the use of DMSO facili-
tates dissolution of the sugar, it can impact negatively on
the enzyme activity although this can be overcome by using
higher concentrations of the acyl donor and/or by employ-
ing, for example, vinyl palmitate as a reactant and thereby
rendering esterification irreversible. This last tactic signifi-
cantly broadens the applicability of such enzymes across a
range of reaction and substrate types.
The capacities of ionic liquids (ILs) to solubilize both
polar and non-polar species provides unique possibilities for
simultaneously dissolving both sucrose and fatty acid esters
and thereby facilitating reaction between them. Given their
green credentials(Otomo et al. 2009; Anastas and Warner
1998), thermal stabilities, salt-like properties in the liquid
state (and that often occurs at room temperature) (Welton
1999), good specific conductivities and negligible vapor pres-
sures, ILs have been exploited in a multitude of settings
(Johnson 2007). They are noted for their abilities to dissolve
biomass components (Singh 2019) and have been employed
as solvents in reactions involving lipases (Itoh 2017) and so
exploited in the preparation 6-Oand 6a-O-linoleyl-a-D-mal-
tose (Fischer et al. 2013).
Figure 4. (a) Lipase esterification of carbohydrates and hydrolysis of carbohydrate esters (Neta, Teixeira, and Rodrigues 2015); and (b) the bi-bi ping-pongtranses-
terification mechanism (Reyes-Duarte et al. 2005).
6 Y. TENG ET AL.
There have been various reports of the enzyme-catalyzed
synthesis of sugar fatty acid esters in ionic liquids (Yang
and Huang 2012). However, only a few of these have
focused on sucrose ester formation, probably because of the
low solubility of sucrose in conventional ILs and/or the
enzyme denaturing properties of such solvents (Shi, Li, and
Chu 2011). As part of efforts to circumvent these difficulties,
the Sheldon group (Liu et al. 2005) established that the solu-
bility of sucrose in [BMIm][dca] is 195 g/L at 25C and
282 g/L at 60 C. Furthermore, they revealed that this IL can
be used in the Novozym 435-catalyzed acylation of sucrose
by dodecanoic acid at 55 C. In related work, Huang et al.
studied the enzymatic preparation of 6-O-lauroylsucrose by
reaction of sucrose with lauric acid vinyl ester and con-
cluded that by deploying Lipozyme TL IM in [BMIM]BF
4
the target ester could be obtained in 50% yield (Huang et al.
2010b). Shao et al. (2018) reported a synthesis of the same
sucrose mono-ester in a 1.5:1 v/v mixture of
[3CIM(EO)][NTf
2
] and 2-methyl-2-butanol and so affording
the product in 66% yield, a pleasing outcome attributed to
the capacity of this IL to maintain the high activity of the
lipase and to enhance the solubilities of the substrates.
Supercritical CO
2
has also been exploited as a solvent in
various lipase-catalyzed biotransformations (Knez 2018). So,
for example, an immobilized lipase from Candida antarctica
B(EC 3.1.1.3) was used in supercritical CO
2
at 10 MPa to
produce sucrose palmitate and sucrose laurate (Habulin,
Sabeder, and Knez 2008). The addition of molecular sieves
to that reaction affording sucrose laurate from equimolar
amounts of the fatty acid and sucrose at 60 C resulted in a
74% conversion. The non-toxic and non-flammable nature
of CO
2
means these types of reactions are considered amen-
able to the industrial preparation of food additives.
Yields and selectivity
Recent reports have revealed that the regioselective produc-
tion of sucrose esters can be effected in 4090% yield using
the lipases derived from Bacillus thermoproteolyticus,
Candida antarctica,T. lanuginosus,Bacillus pseudofirmus
AL-89 and Bacillus subtilis. However, in order to realize
these outcomes acylating agents such as vinyl esters and
anhydrides are required. Interestingly, essentially quantita-
tive yields of a product sucrose ester were achieved using
2,2,2-trifluoroethyl butyrate as the acyl donor in conjunction
with the Bacillus protease Bioenzyme 240. When pyridine
was used as the solvent this combination of reactants pro-
duced 10-O-butyryl sucrose (Patil, Rethwisch, and Dordick
1991). Lipases were less effective. Significantly, lipases
derived from Rhizomucor miehei have been used to effect
the high-yielding formation of sugar mono- and di-esters
with simple carboxylic acids serving as the acyl donors
(Dang, Obiri, and Hayes 2005; Ye, Pyo, and Hayes 2010).
The selective acylation at the 6-Oand 60-Opositions of
sucrose is often observed when lipases are employed for this
purpose (Shi, Li, and Chu 2011) although other less-com-
mon enzymes allow for alternative outcomes. Plous group
reported the preparation of 6-O-lauroylsucrose and 6-O-pal-
mitoylsucrose (Ferrer et al. 1999) through acylation of
sucrose using mixed-solvent combinations and lipases
derived from various species but most notably an immobi-
lized form obtained from Humicola lanuginosa Candida
rugosa lipase (CRL) has also been used in transesterification
studies for the formation of sucrose-6-O-acetate with various
reaction parameters, including temperature, concentration
and pH, being examined and so leading to the identification
of an effective two-phase system comprising 2-butanol and
Tris-HCl buffer at an initial pH of 8.60 and a temperature
of 47 C. Under such conditions the target mono-acetate
was obtained in 57% yield (Zhong et al. 2013).
Reaction parameters
The Hayesgroup has reported detailed studies on the use
of the immobilized lipase derived from Rhizomucor miehei
(viz. Lipozyme IM ex. Novozymes) for preparing sucrose
esters and by such means a 9:1 mixture of mono- and di-
esters was obtained in 8093% combined yield when oleic
acid was employed as the acyl donor (Dang, Obiri, and
Hayes 2005). Notably, only small quantities of solvent (tert-
butanol) were required during the early stages of the
reaction and after a 25% conversion had been realized the
product sucrose ester served as the reaction medium. The
same enzyme and protocols were used for the esterification
of sucrose and fructose with oleic acid. The reaction took
6 days and led to a liquid phase consisting of an 11.5:1 mix-
ture of mono- and di-esters in 88% combined yield (Ye,
Pyo, and Hayes 2010). Similarly high yields were realized
using acetone or methyl ethyl ketone (butanone) instead of
tert-butanol or, alternatively, using carefully defined mix-
tures of oleic acid and fructose or sucrose.
Both Rhizomucor miehei and Candida Antarctica-derived
lipases have been used in esterifications under conditions
involving high-speed homogenization and high-intensity
ultrasonication of the reaction mixtures. These conditions
were found to be effective in reducing the size of sucrose
particles in a solvent-free synthesis of sucrose oleate (Ye,
Hayes, and Burton 2014). It has also been observed that
pre-incubation of C. Antarctica lipase B (CALB) in oleic
acid at 60 C for 24 h enhances the initial rate of enzymatic
esterification. Moreover, the synthesis of sucrose esters using
either acrylic resin or chitosan immobilized CALB in con-
junction with oleic acid, sodium sulfate and ethanol, deliv-
ered product yields of 56 and 55%, respectively (Neta
et al. 2012).
A lipase derived from T. lanuginosus and its immobilized
form, Lipozyme TL IM, have also been used to produce 6-
O-lauryl sucrose from vinyl laurate. Thus, Ferrer et al.
(2002b) prepared a silica-granulated lipase (particle size
around 0.31 mm) obtained from T. lanuginosus, and this so
immobilized enzyme allowed for the synthesis of sucrose
mono-laurate in >95% yield when a mixture 4:1 v/v tert-
amyl alcohol and DMSO was deployed as the solvent system
at 40 C and sucrose added to the reaction mixture in por-
tions. The capacity to recycle this enzyme is highlighted by
its ability to continue to effect this reaction in 80% yield
over 20 cycles, each of which lasts ca. 6 h (Ferrer et al.
2005). The regioselective synthesis of 6-O-palmitoylsucrose
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 7
catalyzed by Lipozyme TL IM in a continuous flow micro-
reactor has been reported (Du and Luo 2012) and involves
using silicone tubing (inner diameter 2.0 mm) with a Y-
shaped junction at the start point that is packed with silica
spheres (0.30.9 mm) carrying surface-adsorbed enzyme. To
carry out the reaction, solutions of sucrose (in 2-methyl-2-
butanol and DMSO) and vinyl palmitate (in 2-methyl-2-
butanol) were injected into the junction for mixing then
flushed through the microreactor. At 40 C (and using a
flow rate of 20.8 lL/min that resulted in a residency time of
0.5 h), a high conversion (95%) of sucrose into the corre-
sponding 6-O-acyl mono-ester was realized. This outcome is
attributed to the favorable surface area-to-volume ratio of
the Lipozyme TL IM presenting on the silica particles.
Proteases
While most reports on the enzymatic preparation of sucrose
esters involve the use of lipases, deploying proteases
(Rawlings et al. 2018) for the same purpose has also been
explored in the ongoing efforts to achieve ever higher regio-
selectivities and yields within useful timeframes. An attract-
ive feature of proteases is their relative stabilities and good
activities when used in conjunction with organic solvents
such as DMF and DMSO (Plou et al. 2002). In general, pro-
teases selectively acylate at the 10-Oposition of sucrose in
such solvents as well as in pyridine (Riva et al. 1988). Thus,
an examination of the effectiveness of subtilisin conducted
by Riva et al. (1988) established that this enzyme is catalytic-
ally active in anhydrous DMF and allows for the production
of 10-O-monobutyrylsucrose in 85% yield and with 90%
selectivity. Using a stabilized, insoluble and commercially
available polymer of substilisin called ChiroCLEC-BL
Linhardt et al. were able to produce a range of 10-O-acyl
sucrose esters that were accompanied by minor amounts of
the corresponding 10,6-di-O-acylated (di-ester) derivatives
(Polat, Bazin, and Linhardt 1997). The most effective reac-
tion conditions employed dry pyridine as the solvent and
vinyl esters as the acyl donors. More recently, Protex 6 L, a
commercially available and alkaline protease derived from
Bacillus licheniformis, was used to synthesize sucrose mono-
laurate (Wang et al. 2012). In this process, vinyl laurate dis-
solved in a 15:1:4 v/v/v mixture of tert-amyl alcohol/DMSO/
water was treated with fine particles of sucrose (0.2 mm).
By such means 98% of the sucrose was converted, regiose-
lectively, into 10-O-lauroylsucrose after 9 h and so establish-
ing that Protex 6 L is superior to many commonly used
enzymes including the immobilized lipase Novozym 435.
The alkaline soda lakes of Ethiopia have proven to be
interesting sources of proteases including one derived from
Bacillus pseudofirmus AL-89 that functions within the pH
range 710 in the presence of 7.5% v/v water (Pedersen
et al. 2003). Under such conditions vinlyl laurate can be
used to effect sucrose esterification at the 2-position rather
than its 10or 6-counterparts. Another 2-O-selective acetyl-
ation of sucrose is achieved using the neutral metallopro-
tease thermolysin (E.C. 3.4.24.27) in DMSO. This catalytic
process relies, as confirmed by EDTA binding and other
studies, on the presence of zinc in the active site (Pedersen
et al. 2002; Lie, Meyer, and Pedersen 2014).
Liaos group has reported regioselective acetylation at the
40-Oposition of sucrose using a purified serine protease
derived from Serratia sp. SYBC H and thus providing
sucrose-40-Omono-acetate in more than 90% yield (Li et al.
2011a). In contrast, when a crude alkaline protease obtained
from the same source was used then exhaustive and high
yielding peracetylation of sucrose was observed (Li
et al. 2011b).
Other enzymes
A 2016 study focused on the preparation of sucrose 6-O
acetate employed Aspergillus oryzae fructosyltransferase
(E.C. 2.4.1.9 or, as recommended by BRENDA, inulosucrase)
(BRENDA 2019a) that was originally isolated from a sample
of root soil associated with the growth of sugarcane (Wei
et al. 2016). This enzyme, which has been shown to transfer
a fructosyl unit from one sucrose residue to another (Wei
et al. 2014), was deployed in various IL/buffer combinations
with sucrose and glucose 6-Oacetate as substrates. The best
medium for the reaction proved to be a 1:4 v/v
[Dmim][PF
6
]/phosphate buffer system and under optimal
conditions 88% conversion of glucose 6-Oacetate was
observed and the product sucrose 6-Oacetate obtained in
77% yield. In 2017, a novel, chemoenzymatic route to a new
class of sucrose esters was reported (Possiel, B
auerle, and
Seibel 2017). The enzyme used in this case was Bacillus sub-
tilis levansucrase (EC 2.4.1.10, a member of the glycoside
hydrolases family of 68 enzymes) that is known to hydrolyze
sucrose as well as effect transfer of the fructosyl moiety of
sucrose (Ortiz-Soto et al. 2017). By such means, the glucose
residue of sucrose was substituted for various D-glucuronic
or D-galacturonic acid ester residues to form the corre-
sponding b-D-fructofuranosyl-(2,1)-a-D-glucuronic or galac-
turonic acid esters.
Table 2. Commercial enzymes used in the preparation of sucrose esters.
Enzyme Other Names
Lipase (Thermomyces lanuginosus, previously Humicola lanuginosa) (Ferrer
et al. 2005)
Lipozyme
V
R
TL IM (Shao et al. 2018)
Lipase (Candida Antarctic B) Novozym 435
V
R
(Ye, Hayes, and Burton 2014)
Lipase (Rhizomucor miehei) Lipozyme IM
V
R
(Dang, Obiri, and Hayes 2005; Ye, Pyo, and Hayes 2010)
Lipase (Candida rugosa) CRL (de Maria et al. 2006)
Serine protease subtilisin (Bacillus licheniformis) ChiroCLEC-BL (Polat, Bazin, and Linhardt 1997); Protex 6L (Wang et al. 2012);
Subtilisin A, Subtilopeptidase A, Novozymes Alcalase
V
R
(Rawlings et al.
2018); Subtilisin Carlsberg (Delange and Smith1968)
8 Y. TENG ET AL.
Summary
Certain of the commercially available enzymes used for the
preparation of sucrose esters are listed in Table 2 and their
selectivities are summarized in Figure 5. Further, and as out-
lined above and emphasized in Table 3, enzymes can pro-
vide greener and more selective methods for obtaining
sucrose esters, including ones that are often inaccessible by
other means. Nevertheless, much remains to be done within
this promising area of sucrose mono-ester synthesis, both
for the purposes of making customizedand essentially
pure forms of such systems in boutiquequantities as well
as in the production of such systems at commercial scales
for the food industry. The use of, inter alia, protein
engineering, directed evolution and gene shuffling techni-
ques for the production of dedicated enzymes capable of
meeting industry demands for superior sucrose-ester pro-
duction processes will certainly be a focus of ongoing efforts
in this area.
Sucrose esters as emulsifiers
Sucrose esters are of great interest as surfactants. For more
than fifty years the food industry has been developing meth-
ods for preparing various sucrose esters with a tunable
range of surface activities (Tucker and Martin 1958; Roshdy,
Horst, and Herbert 1967). As a result, a large number of
Figure 5. The regioselectivities exhibited by various enzymes available for the formation of sucrose esters and certain derivatives.
Table 3. Brief comparison of various methods available for the synthesis of sucrose esters.
Synthesis Method Advantages Disadvantages
chemical preparation methods in general catalyst cost is usually low;
some economical processes have been
industrialized
are generally less selective and so yield a mixture
of esters;
complicated product purification;
generation of waste streams
transesterification in solvents (typically DMF) providing >70% of monoester;
preferred industrially
solvents are often toxic and difficult to remove
solvent-free inexpensive raw material and catalyst cost;
can be industrialized
forms a mixture of esters;
degradation of sucrose leads to darkening
with acid chlorides enhanced proportions of mono-esters strict control of water content required;
high catalyst cost;
unfavorable in industry
with fatty acid vinyl esters (also applicable
with enzymes)
the irreversible reaction affords high yields and
high mono-ester content;
often preferred in lab settings
high cost of acylation reagent and hence
unfavorable in industry
with other acylation agents (e.g., acid anhydrides
and enol esters)
can achieve high selectivity complicated reaction procedures makes it
unfavorable in industry
under Mitsunobu conditions preparation of functional esters (e.g., gallic acid
ester) in lab
complicated reaction procedures make it
unfavorable in industry
enzymatic preparation methods in general high selectivity;
green processes and purer products
high catalyst cost;
industrialization is challenging
in organic solvents facilitates dissolution of sugar organic solvents can deactivate enzymes
in ionic liquids can dissolve both substrates and thus facilitate
the reaction
low solubility of sucrose in conventional ionic
liquids;
enzyme denaturing
in supercritical CO
2
solvent is non-toxic and non-flammable high production cost
reduced solvent and solvent-free system avoids massive use of solvents small sucrose particles are needed;
homogenization or ultrasonication is often
needed to promote reaction
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 9
sucrose esters is now available commercially for satisfying
diverse needs in a wide range of areas. Various parameters
have been used to characterize and calibrate their properties
as surfactants (Table 4) and so assisting with their commer-
cial success (Griffin 1949; Griffin 1954; Ren and Lamsal
2017; Pearce and Kinsella 1978; McClements 2015).
The emulsifying capacity of surfactants can be very
clearly determined/demonstrated by measuring the particle/
droplet sizes and their distributions associated with the
emulsions formed using them. In order to examine droplet
size, various devices, including laser diffraction particle-size
analyzers, can be employed and a plot of continuous size
distribution is then generated by the software packages nor-
mally associated with such instruments. In general, a smaller
mean value and narrow standard deviation of the distribu-
tion signifies a better emulsifying capacity (Walstra 2002).
McClements (2015) suggested that droplet size will influence
a range of properties of the emulsions including stability,
optical behaviors, rheology and even sensory attributes (Ren
and Lamsal 2017; Pearce and Kinsella 1978).
HLB values and their utility
Sucrose esters have unique surface-active properties arising
from their amphiphilic nature that itself derives from the
presence, within the molecular framework, of the hydro-
philic hydroxyl groups associated with the sucrose moiety
and hydrophobic characteristics associated with the fatty
acyl side-chain. The hydrophilicity or lipophilicity of a sur-
factant is often measured by the so-called hydrophilic-lipo-
philic balance (HLB) that ranges in value from 0
(completely lipophilic/hydrophobic) to 20 (completely
hydrophilic/lipophobic) (Griffin 1949; Griffin 1954). In the
case of sucrose esters, a wide range of HLB values can be
achieved by controlling the type of associated fatty acids and
the degree of esterification of the carbohydrate framework
(Plou et al. 2002). As would be expected, longer acyl chains
or higher degrees of esterification will increase the
hydrophobicity and thereby decrease the HLB value. In
practice, the HLB values of commercial sucrose esters range
between 1 and 16 (Mitsubishi Chemicals 2019; Sisterna
2019a). It is also possible to blend sucrose esters with other
surface active compounds (often glycerides) to meet differ-
ent needs (Gupta, James, and Smith 1983). As noted by Ye
and Hayes (2014), sucrose esters with HLB values 5 are
used in chocolates to form water-in-oil emulsions while
those with intermediate HLB values (810) are best used in,
for example, chewing gum for their more balanced lipophi-
licity and hydrophilicity. Those esters with HLB values 11
are widely used in ice cream and cake batter to form oil-in-
water (an opposed to water-in-oil) emulsions. The type of
fatty acid residue involved also impacts of the utility of the
sucrose esters. For example, when the HLB value of the sur-
factant is between 11 and 16, as seen in sucrose esters with
12, 14, 16, 18 and 18:1 carbon-containing acyl groups (viz.
in laurate, myristate, palmitate, stearate and oleate esters
derivatives), these are used as detergents, in ice cream, in
milk beverages, in cake batter, and in sauces or dressings,
respectively (Ye and Hayes 2014). A great deal of informa-
tion about the surface-active properties and applications of
sucrose esters is provided in the scientific literature (Otomo
et al. 2009; Ye and Hayes 2014) and/or on manufacturers
websites (Mitsubishi Chemicals 2019; Sisterna 2019a).
Various new studies relating to the HLB values of sucrose
have been reported. So, for example, Jiang et al. (2019)
investigated the competitive interfacial adsorption of com-
mercial emulsifiers such as Mitsubishis lipophilic sucrose
stearate (S-370, HLB ¼3) and its more hydrophilic counter-
part (S-1670, HLB ¼16). They found that, overall, the for-
mer was superior in producing well-textured frozen aerated
emulsions. These same authors claimed that lipophilic emul-
sifiers accelerated crystallization of fat, promoted interfacial
heterogeneous nucleation as well as partial coalescence of fat
droplets, enabled strong aeration activity, and allowed a
short melting starting time. Sucrose ester S-1170 (HLB ¼
11) was also added (in up to 6% w/w) into a high amylose
starch-based wood adhesive so as to improve bonding/shear
Table 4. Parameters commonly used to characterize the emulsifying effects of surfactants.
Parameter and formula Definitions of variables
Hydrophilic-lipophilic balance (HLB) value (Zhang et al. 2014):
HLB ¼20 Mh
M
M
h
:molar mass of the hydrophilic moiety.
M:molar mass of the surfactant molecule.
Foamability (Zhang et al. 2014):
FA ¼HTH0
H0100%
H
0
:height of surfactant solution before foam formation.
H
T
:total height of foam and solution measured immediately after foam formation.
Foaming stability (Zhang et al. 2014):
FS ¼Ht
Hi100%
H
i
:foam height measured immediately after foam formation.
H
t
:foam height after the sample was rested for some time.
Emulsifying ability (Zhang et al. 2014):
EA ¼H1
H0100%
H
0
:height of surfactant solution and an oil phase (e.g., equal volume of soybean oil)
H
1
:height of the emulsion layer measured immediately after homogenization.
Emulsion stability (Zhang et al. 2014):
ES ¼H2
H1100%
H
1
:height of the emulsion layer measured immediately after homogenization.
H
2
:height of the emulsion layer after the emulsion was rested for some time.
Emulsion stability index (Ren and Lamsal 2017; Pearce and Kinsella 1978):
ESI ¼A0t
A0At100%
A
0
:initial absorbance obtained immediately after homogenization and dilution.
A
t
:absorbance obtained after the test sample was rested for certain time.
t: time that the test sample was rested (e.g., 20 min).
Size distribution (McClements 2015):
Sn¼Ð1
0xnFx
ðÞ
dx Pi¼1nixn
i
xab ¼Sa
Sb

1=ðabÞ
cn¼SnSnþ2
S2
nþ1
1

1=2
S
n
: the nth moment of the distribution.
x
ab
: the mean particle size.
a,b: integers (usually between 0 and 6).
c
n
: relative standard deviation weighted with the nth power of x.
n: the nth moment of the distribution.
n
i
: the number of droplets in a size-class.
10 Y. TENG ET AL.
strength (in both dry and wet states), mobility, storage and
thermal stability as well as rheological and degradation
resistant properties. At the micro-level, the addition of this
sucrose ester was found to hinder the aggregation of latex
particles as well as their dispersion (Zia-Ud-Din et al. 2017;
Zia-Ud-Din et al. 2019). Sucrose esters were also added to
tobacco shred so as to improve their humectant and mois-
ture resistant properties. Similarly, the hydrophobic portion
of the sucrose ester S-5 (HLB ¼5), as supplied by the
Zhejiang Synose Tech Co., Ltd., was shown to play an
important role in reducing the interaction of tobacco shred
with water (Lin et al. 2020). Likewise, sucrose esters,
together with other small molecular emulsifiers, were found
to reduce the degradation of starch in food products like
pancake mixtures (Yamashita et al. 2020). This reduction
was attributed, in the main, to emulsifier-starch complex
formation and so reducing the leaching of starch molecules
from product granules. The degree of such complexation
was, in turn, influenced by the HLB values of the emulsi-
fiers involved.
The development of synthetic methods (in particular
enzymatic ones) has enabled the preparation of high purity
sucrose mono-esters and these have allowed for in-depth
comparative studies of their surface-active properties (Ferrer
et al. 2002a). So, Zhang et al. (2014) have reported that a
longer acyl chain (C812) results in a reduction of HLB
(14.5 to 13.1) and CMC (0.780.45 mM) values while the
associated
cCMC
values increase (32.3634.54 m/Nm). By
comparing fatty acid sugar mono-esters incorporating
C1218-containing acyl chains, it was found that those
involving the longer ones decrease the CMC values while
for those embodying the same length of acyl chain the order
of CMC values was: sucrose >maltose >leucrose >
maltotriose (Ferrer et al. 2002a). All in all, most studies of
the surface properties of sucrose esters reveal that they are
good surfactants with a wide range of practical applications.
Foaming properties
Much confectionery, including ice cream, mousses and
marshmallow, use food emulsifiers to provide a satisfying
texture, attractive appearance and long shelf-lives. As such,
the so-called foamability of these emulsifiers as well as the
stability of the derived foams (foam stability) are of vital
importance (Hutzler et al. 2011). Surfactants stabilize the
foams by preventing bubbles from coalescing (Belhaij and
Al-Mahdy 2015) so foamability, which is defined as the sur-
factantscapacity to form foams, and foam stability (a metric
derived by determining variations in foam height or volume
as a function of time), are important metrics used to define
the foaming properties of food emulsifiers (Husband
et al. 1998).
At the early stage, Husband and coworkers demonstrated
that a 4:1 molar ratio of sucrose mono- and di-esters can
display superior foaming properties as a result of interac-
tions between these constituents (Husband et al. 1998). It is
also widely recognized that the stability of foams derived
from sucrose octaacetate decreases on standing (Petkova
et al. 2017). Tual et al. (2006) have suggested that the desta-
bilization of fat droplet interfacial layers in dairy products
can be caused by the presence of sucrose esters meaning
that optimal concentrations of these must be established so
as to produce the most stable and high-quality dairy foams.
Furthermore, when the concentration of the sucrose ester is
too low to support foam formation then adding small
amounts (often just 0.050 wt.%) of b-lactoglobulin can
retrieve matters (Garofalakis and Murray 2001). Although
aqueous foams are more common than non-aqueous ones,
more research on the latter has been reported due to their
great potential (Fameau and Saint-Jalmes 2017). For
example, the edible emulsifiers made from sucrose esters
and lecithin contribute to the formation of stable oil foams
involving high air volume fractions and thus showing inter-
esting rheological and thermoresponsive properties likely to
be of value in generating food products possessing novel
textures (Patel 2017).
Emulsifying capacity
Sucrose esters are expected to provide emulsifiers suitable
for deployment in various food products and cosmetics, not
least because of their broad range of HLB values, their non-
toxic nature and their degradation to readily digestible
sucrose and fatty acids. As such, several methodologies and
metrics have been developed so as to quantitate their emul-
sifying capacities (Otomo et al. 2009). For example,
Ariyaprakai and coworkers have reported that sucrose esters
could be employed as a promising replacement for Tweens
in coconut milk emulsions since such substitutions provide
more thermally stable systems (Ariyaprakai, Limpachoti, and
Pradipasena 2013). This group also reported that edible oil-
in-water emulsions derived from sucrose esters are particu-
larly stable toward freezing then thawing (Ariyaprakai and
Tananuwong 2015). Furthermore, Zhao, Chen, and Wu
(2018) demonstrated that the addition of carefully selected
cryogelling polysaccharides such as alginate can enhance the
stability of the associated emulsions toward freeze-thaw
cycles. Similarly, mixtures of polyglycerol monostearate
(PGE) and sucrose esters have proven to be useful in the
preparation of flavor oil emulsions (Ariyaprakai 2016).
Other research has revealed that sub-micron-sized emulsions
can be formed from sucrose esters via a so-called spontan-
eous emulsification process instead of employing higher-
energy homogenization techniques (Ariyaprakai, Hu, and
Tran 2019) and that, under simulated gastrointestinal condi-
tions, sucrose ester-derived emulsions undergo coalescence
while Tween80-based ones do not (Verkempinck
et al. 2018).
Biological activities
While sucrose esters are considered to be safe for humans, a
multitude of studies have reported their inhibitory effects on
microbes, tumor cells and viruses as well as revealing their
insecticidal potential. Such properties are now discussed.
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 11
Antimicrobial activity
Since sucrose esters have been widely used in food products,
their antimicrobial activities (including antifungal and anti-
bacterial effects) have been also been studied extensively as
highlighted by the data presented in Table 5. Most early
research in this area used commercially-derived sucrose
esters but in recent years studies have been focused on more
novel and enzymatically-derived systems (Ferrer et al. 2005).
It has been known for sometime that detergent mole-
cules, such as polyoxyethyleneglycol, can bind to bacterial
membranes and so destroying them (Le Maire, Champeil,
and Møller 2000), and it is very likely that sucrose esters
exert their antibacterial effects by the same basic mechanism.
Indeed, as revealed by Shao et al. (2018), the antimicrobial
activity (particularly against Gram-positive bacteria) of
sucrose monolaurate is initiated through disruption of the
integrity of the bacterial cell membrane. This triggers col-
lapse of the cytoplasmic membrane, loss/leakage of intracel-
lular enzymes and the release of K
þ
from the cytosol,
disruption of the subcellular localization of proteins and,
thereby, bacterial inactivation.
Thanks to their antimicrobial properties, sucrose esters
are considered to have a promising future as non-toxic food
preservatives while also possessing unique emulsifying and/
or foaming properties (Habulin,
Sabeder, and Knez 2008).
Yang et al. (2003) suggested that the use of sucrose mono-
esters derived from lauric, myristic and palmitic acids in
salad dressing will also inhibit the growth of microorganisms
involved in spoilage. In fact, sucrose esters even show
inhibitory effects on the growth of spores, prompting
Shearer et al. (2000) to suggest using sucrose laurate for
food applications where high-temperature processing is
undesirable. Another interesting use of sucrose esters is in
oral-care products like toothpaste. It has been reported that
6-O-lauroylsucrose completely inhibits the growth of the
oral bacterium Streptococcus sobrinus and so suggesting that
such compounds could be included in oral-hygiene products
in order to disrupt plaque formation and thereby preventing
(or at least reducing) dental caries (Devulapalle et al. 2004).
Given their increasingly broad deployment, the possible
emergence of microorganisms resistant to sucrose esters has
been investigated. As Sugimoto et al. (1998) discovered,
while some commercial sucrose mono-esters inhibit the
development of Bacillus cereus spores as well as the vegeta-
tive growth of cells, the germinated spores and growing cells
can release esterases that promote ester degradation and so
aiding their capacity to regenerate (viz. become resistant).
Antitumor activities
In 1971, Kato et al. showed that the water-soluble sucrose
esters incorporating laurate, myristate, palmitate and linole-
ate residues exhibit strong in vitro antitumor activity against
Ehrlich ascites tumors in mice. Subsequently, the Nishikawa
group reported a series of studies on the chemical and bio-
chemical properties of synthetic and commercially available
carbohydrate esters including sucrose esters. The group first
revealed that sucrose mono-esters of myristic acid
(Nishikawa et al. 1976a) as well as elaidic and oleic acids
(Nishikawa et al. 1976b) exerted marked inhibitory activities
against Ehrlich ascites carcinoma in mice. Subsequently,
they found that sucrose mono-esters of palmitic, stearic,
hydnocarpic and ricinoleic acids exerted antitumor effects
against this cancer as well (Nishikawa et al. 1977a). These
studies have been extended to commercial sucrose esters as
well as other tumor cell lines and so revealing that mono-
ester-rich systems exhibit pronounced antitumor effects
(Nishikawa et al. 1977b; Ikekawa et al. 1979). A consider-
ation of the possible mechanisms of actions of these esters
as cytotoxic agents suggested that they might not just act as
detergents (and so disrupting tumor cell membranes) or as
prodrugs for the associated fatty acids (Nishikawa et al.
1976a) but also have inherent antitumor properties them-
selves (Nishikawa et al. 1977a). Such a possibility is sup-
ported by the observation that as the HLB value of the
sucrose monoester increased the associated ID
50
value
decreased (Ikekawa et al. 1979).
Interestingly, plant-derived sucrose esters can also exhibit
antitumor activities. Thus, in 2017, Fang et al. reported the
isolation of eight isovaleryl sucrose esters (named ainslosides
AH) from Ainsliaea yunnanensis Franch, and discovered
that congener B (Figure 6a) exerted notable cytotoxic effects
against the A549 (lung adenocarcinoma) cell line with an
IC
50
of 3.3 lM. This compound was found to arrest the cell
cycle at the G
0
/G
1
phase and induce cell apoptosis with the
latter effect arising from a decrease in mitochondrial mem-
brane potential and an attendant growth in the levels of
growth of reactive oxygen species. In 2018, Petrova et al.
synthesized a library of sugar esters with C35 unsaturated
fatty acyl chains and found that 6-O-methacryloyl sucrose as
well as 10,2,3,30,4,40,60-hepta-O-acetyl-6-O-methacryloyl
sucrose (Figure 6b) were the most active against a range of
food contaminating and clinically notable pathogens (MIC:
0.241.40 lM). The later ester also showed good antifungal
activity (MIC: 0.281.10 lM). Synthetic, non-fatty-acid
sucrose esters possessing antitumor activities have also been
identified. For example, Bai (2009) prepared a unique
sucrose selenite (Figure 6c) showing promise in treating can-
cers. This compound has a selenium content of 17.4% which
is significantly higher than that of conventional selenium-
containing nutrients and anticancer drugs (ca. 1.2%). It
exhibited excellent antimutagenic bioactivity as evidenced by
an inhibition rate of 97.4% when applied in a standard assay
at a loading of 500 lg/dish.
Sucrose fatty acids esters have also been used in drug
delivery systems (Abdel-Mageed et al. 2012; Guan, Chen,
and Zhong 2019). For example, Guan, Chen, and Zhong
(2019) recently reported that a sucrose fatty acid ester pur-
chased from the Tokyo Chemical Industry Co. in Japan
(Product No. S0112) could be used to nanoencapsulate caf-
feic acid phenethyl ester (CAPE), a natural product possess-
ing anticancer activities but low water solubility. Such
encapsulation is expected to enhance the capacity to deploy
CAPE in the treatment of certain colon and breast cancers.
Of course, the possible synergistic effects associated with the
12 Y. TENG ET AL.
Table 5. Summary of the literature reporting the antimicrobial activities of sucrose fatty esters.
Sucrose ester Active Against Comment Ref.
6-O-lauroylsucrose L. monocytogenes,B. subtilis,S.
aureus, and E. coli.
Indicates significant antimicrobial
activity against these bacteria,
particularly Gram-positive bacteria.
Shao et al. 2018
Sucrose monolaurate
(6-O-lauroylsucrose, from Sigma-
Aldrich Co.)
Gram-positive bacteria Exerted bacteriostatic and bactericidal
effects against these bacteria.
Park et al. 2018
Physakengoses
(natural sucrose esters isolated
from Physalis alkekengi
var. franchetii)
Staphylococcus aureus,Bacillus subtilis,
Pseudomonas aeruginosa, and
Escherichia coli.
Five out of seven natural sucrose
esters isolated displayed potent
antibacterial activity, and long
chain fatty acid esters attached to
sucrose were essential for
this activity.
Zhang et al. 2017
Sucrose monocaprate
(home-made)
Bacillus cereus,Bacillus subtilis,
Staphylococcus aureus,Escherichia
coli, and Salmonella typhimurium.
Displayed the strongest antibacterial
activity against these bacteria,
particularly Gram-positive bacteria.
Zhao et al. 2015
Octa-O-acetylsucrose 17 Microorganisms:
Gram-positive and Gram-negative
bacteria, yeasts, and fungi
Inhibited the growth of fungi
Penicillium sp., Rhizopus sp. and
Fusarium moniliforme, and yeasts
Candida albicans. Did not inhibit
Gram-positive and
negative bacteria.
Petkova et al. 2017
Sucrose esters with C8, C10, C12, C14
and C16 fatty acids
(Mitsubishi-Kagaku Foods
Co., Japan)
Bacillus species: B. cereus,B. subtilis,
B. megaterium, and B. coagulans
Sucrose esters displayed bactericidal
effects on these species at pH 6.0,
but only showed an antibacterial
effect on B. coagulans at pH 8.0
(better at 37 C than 50 C).
Nakayama et al. 2015
Sucrose monolaurate
(from Sisterna, the Netherlands)
Gram-positive and Gram-
negative bacteria
Most effective against the growth of
Gram-positive bacteria (inhibited
Listeria and greatly inhibited
S. suis).
Wagh et al. 2012
Sucrose monolaurate
(from AppliChem Inc.,
Darmstadt, Germany)
Escherichia coli O157:H7 The efficacy of chlorine (from NaOCl)
sanitization was significantly
improved by sucrose monolaurate
which enables inactivation or
removal of the microorganism.
Xiao et al. 2011
Commercial sucrose ester
(sucrose palmitate, P-1670, from
Mitsubishi-Kagaku Food Co., 80%
mono ester), or enzymatically
synthesized sucrose laurate (mono
and diesters)
Gram-positive bacterium B. cereus
(ATCC 11778)
At 9.375 mg/mL, commercial and
enzymatically synthesized sucrose
laurate inhibited 99.8% and 94.0%
of the growth of B. cereus after
52 h.
Mono-esters inhibit B. cereus to
higher extent than di-esters.
Habulin,
Sabeder, and Knez 2008
6-O-lauroylsucrose (2 mg/mL)
6,10-di-O-lauroylsucrose (1 mg/mL)
6,60-di-O-lauroylsucrose
(0.25 mg/mL)
Bacillus sp.,Pseudomonas fluorescens,
Staphylococcus aureus,Escherichia
coli, and Pichia jadinii
Inhibited the growth of Bacillus sp.,E.
coli and L. plantarum. But the
antimicrobial activity of these
diesters was negligible as a result
of their limited solubility.
Ferrer et al. 2005
6-O-lauroylsucrose Streptococcus sobrinus In media supplemented with
2 mg/mL of the sucrose ester,
bacterial growth was
not observed.
Devulapalle et al. 2004
Sucrose monoesters of lauric,
myristic, palmitic, and stearic acids.
Zygosaccharomyces bailii, and
Lactobacillus fructivoran
The growth of Z. bailii was
significantly inhibited by 1%
sucrose monoesters of lauric,
myristic and palmitic acid in salad
dressing, but the inhibition of L.
fructivorans lacks practical value.
Husband et al. 1998
Sucrose laurates, palmitate, and
stearates (Mitsubishi Chemical Co.,
N.Y., U.S.)
Spores of Bacillus sp., Clostridium
sporogenes PA3679, and
Alicyclobacillus sp.
In terms of sporeformer inhibition,
sucrose laurates and palmitate
were more effective
than stearates.
Shearer et al. 2000
Sucrose palmitate (P-1570 and P-
1670)
Sucrose stearate (S-1570 and S-
1670)
(Mitsubishi-Kagaku Foods Co.)
L. monocytogenes,Bacillus cereus
(both cells and spores),
Lactobacillus plantarum, and
Stuphylococcus aureus
Improved the antimicrobial activity of
nisin (an antibacterial peptide
used as a food preservative)
against these strains. Inhibition of
Gram-negative bacteria was
not observed.
Thomas et al. 1998
Sucrose monolaurate (SE12), sucrose
monopalmitate (SE16), and sucrose
Bacillus cereus Inhibit the germination of spores,
with the effectiveness of
laurate >palmitate >stearate, and
Sugimoto et al. 1998
(continued)
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 13
simultaneous deployment of two proven cytotoxic agents in
such settings is worthy of investigation.
Insecticidal activities
A less studied aspect of sucrose esters properties concerns
their insecticidal effects. In 1994, field and green house stud-
ies of Nicotiana gossei Domin plants led to the identification
of an isolate, containing 2,3-di-O-acyl-10-O-acetylsucrose and
2,3-di-O-acyl-10,60-di-O-acetylsucrose, that proved toxic to
sweet potato whiteflies. The side-chains associated these esters
were predominantly derived from 5-methylhexanoic and 5-
methylheptanoic acids (Figure 6d) (Severson et al. 1994).
Subsequent reports described the insecticidal activity of cer-
tain synthetic sugar esters, including sucrose-based ones,
against a range of insects including Bemisia argentifolii (Liu,
Stansly, and Chortyk 1996; Chortyk, Pomonis, and Johnson
1996). Synthetically-derived sucrose esters containing C
6
C
12
aliphatic acid residues (Youngs 1958) and especially dihepta-
noyl, dioctanoyl and dinonanoyl-containing ones, were shown
to be active against whiteflies and aphids (Chortyk 2003)and
field tests of these (to protect various crops) were undertaken.
The prospect of using Nicotiana species for producing sugar
esters as biorationalinsecticides has also been pursued
(Jackson et al. 1998).
McKenzie and Puterka (2004) have studied the insecti-
cidal activities of sucrose octanoate and in so doing found
that nymphal and adult Diaphorina citri (Asian citrus psylla)
as well as mites were both controlled to considerable effect
(>90%) by sucrose octanoate, albeit at rather high concen-
trations (8000 ppm). Given that sucrose octanoate is essen-
tially nontoxic to many beneficial insects, this ester may
prove to be a useful (selective) pesticide in commercial set-
tings. Song et al. (2006), who prepared sucrose octanoate by
transesterification of sucrose with ethyl octanoate under
reduced pressure, have also evaluated the insecticidal activity
of this compound. They reported that the contact toxicity of
sucrose octanoate to first-instar larvae of Lymantria dispar
was 72.5% after 36 h, and the insect reduction rate, after
5 days, of Aphis glycines was above 80% at application rates
of 4 and 8 mg/mL. Considering the emulsifying properties
and other advantages of sucrose esters (such as their bio-
degradability as well as lack of toxicity toward higher ani-
mals and crops), they represent promising leads for the
development of commercial insecticides.
Puterka et al. (2003) studied the structure-activity profiles
of various sugar esters in an effort to determine the origins
of their insecticidal properties and claimed that changing
the fatty acid and sugar residues in such systems resulted in
unpredictable levels of insecticidal activity. Notably, their
HLB values did not correlate with activity. So, for example,
they observed that monoester-rich sucrose octanoate had
higher activities than other esters embodying acyl chains of
similar length (C612) and was toxic to a broader range of
arthropod species. Interestingly, a comparison of the insecti-
cidal activities of sugar esters and soaps led to the conclu-
sion that more than one mechanism-of-action might be
involved with one of these being suffocation caused by
mechanical occlusions of body openings and another being
desiccation of insect cuticles.
Structure-property profiles
In general, the overall surfactant and emulsifying properties
of sugar esters are profoundly influenced by both the length
of the associated fatty acid chain (viz. the hydrophobic tail)
and the nature of the sugar residue (viz. the hydrophilic
head) (Zhang et al. 2015). In the case of sucrose esters,
Table 5. Continued
monostearate (SE18)
(Mitsubishi-Kagaku Foods Co.)
the vegetative growth of cells. But
the germinated spores and
growing cells can release esterases
to decompose these esters,
causing loss of
antimicrobial activity.
Sucrose monocaprate, and
Sucrose monolaurate
Listeria monocytogenes, and
Staphylococcus aureus
The monocaprate (400 lg/mL) did
not inhibit growth, while the
monolaurate exhibited an
inhibitory or lethal effect and this
can be enhanced by EDTA.
Monk, Beuchat, and Hathcox 1996
Sucrose monolaurate Streptococcus mutans NCTC 10449
(an oral bacterium)
Reduced acid production from sugar.
This is due to altering of cell
membrane permeability causing
loss of important metabolites.
Iwami, Schachtele, and Yamada 1995
Six sucrose esters with fatty acid
composition of ca. 70% stearic
acid and 30% palmitic acid. (DKS
International Inc., Tokoyo, Japan)
Aspergillus,Penicillium,
Cladosporium, and Alternaria.
Antimycotic activity was detected
against listed molds. The least
substituted sucrose ester was the
most active, and this inhibitory
activity was not influenced by
variations in pH.
Marshall and Bullerman 1986
Sucrose caprate, caprate, laurate,
myristate and palmitate (Ryoto Co.
Ltd., Tokyo, Japan.)
Vibrio parahaemolyticus The minimum concentrations for
inhibition by sucrose caprylate and
caprate were 40 and 100 lg/mL,
respectively. The rest of the esters
studied were ineffective at
100 lg/mL.
Beuchat 1980
14 Y. TENG ET AL.
precisely how their properties vary as a function of the
length of the fatty acid residue (structure-property profiles)
remains unclear. Zhang et al. (2015) have prepared and
characterized three sucrose medium-chain fatty acid (octa-
noate, decanoate and laurate) mono-esters and found that
increasing hydrophobic chain length enhances the emulsify-
ing potencies of medium-chain fatty acid mono-esters.
However, sucrose mono- and di-esters bearing longer (>12
carbons) chain fatty acids have yet to be properly evaluated
in this regard. As well as the acyl chain length, it can be
expected that increasing the degree of acylation/esterification
leads to a higher hydrophobicity and so decreasing surface
activity. A more important issue is the associated decline in
water solubility, a property that limits applications in the
food industry (Ferrer et al. 2002a). In principle, the degree
of unsaturation of the acyl chain should have limited influ-
ence on the surface properties, as it does not significantly
alter the molar mass of the hydrophilic moiety and, hence,
the HLB. Although the practical applications of sucrose
esters in the food industry are guided, at least in a rudimen-
tary sense, by their hydrophilic-lipophilic balance (HLB) val-
ues (Ye and Hayes 2014), the development of detailed
structure-property profiles are wanting. Once established
these would be expected to aid in the selection of the most
relevant ones for any given practical application(s).
In terms of the bioactivities of sucrose esters, it seems
that mono-esters rather than higher order ones play pivotal
roles, and the influence of the acyl chain length may also be
Figure 6. Selected sucrose esters showing biological activities: a) ainsloside B (Fang et al. 2017); b) sucrose esters of some unsaturated fatty acids (Petrova et al.
2018); c) sucrose selenite (Bai 2009); and d) the natural pesticides 2,3-di-O-acyl-10-O-acetylsucrose and 2,3-di-O-acyl-10,60-di-O-acetylsucrose (the acid residues of
which are mostly derived from 5-methylhexanoic and 5-methylheptanoic acids) (Severson et al. 1994).
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 15
involved. Thus, it has been shown that mono-esters are sig-
nificant antibacterial entities with the corresponding di- and
tri-esters displaying less notable activities possibly because of
their low(er) aqueous solubility. The length of the fatty acid
residue has a notable influence on activity
(C12 >C10 >C8), almost certainly a result of its influence
on the HLB that is involved in the inhibitory effect (Zhang
et al. 2014). In terms of the antitumor activity, the
Nishikawa group has concluded that mono-esters are most
active and that there might often be an inverse correlation
between HLB and the ID
50
values (Nishikawa et al. 1977b;
Ikekawa et al. 1979). Unsaturation within the fatty acyl
chains (Nishikawa et al. 1976a,1976b) has little impact on
activity unless a Michael acceptor is involved (e.g., Petrova
et al. 2018). In terms of insecticidal activity, the impact of
the fatty acid residue is unpredictable although a high
mono-ester content sucrose octanoate exhibits high toxicity
to a broad ranges of species (Puterka et al. 2003). Clearly,
more research is required so as to fully reveal the mecha-
nisms-of-action of such compounds and thereby establishing
guidelines for their practical applications.
In our recent work on homologous series of glucose, mal-
tose, lactose and raffinose esters (Liang et al. 2018a,2018b;
Ma et al. 2018;Lietal.2019) we established that in order to
maintain good foaming properties, fatty acid side chains of
10 or 12 carbons in length are required regardless of the
nature of the sugar residue involved. So, while the incorpor-
ation of longer fatty acid side-chains results in better emulsi-
fying potencies for glucose and raffinose esters, medium
length ones are more effective when bonded to their maltose
and lactose counterparts. Intriguingly, in the case of lactose
esters, the longer the alkyl side chain the greater their cyto-
toxicities. However, determining analogous and unequivocal
structure-property profiles for sucrose esters is complicated
by the mixed nature of most commercially available materials
that can often include seven or more types of fatty acid resi-
dues. Clearly further work in this important area is required.
Conclusion
In recent years, there has been a growing interest in the use
of sucrose esters as emulsifiers because of their excellent
properties, especially as these can be exploited in the food
industry. An associated body of knowledge focused on their
preparation has accumulated as a result and attendant,
detailed studies of their emulsifying properties and various
bioactivities have emerged. While rather conventional chem-
ical methods dominate the industrial production of such sys-
tems, a more recent but largely academic focus has been on
the enzymatic preparation of sucrose esters. There are dis-
tinct advantages associated with such approaches given that
they are generally greener and less energy-demanding proc-
esses but the likely cost of the enzyme required in applying
such protocols in industrial settings remains a challenge.
The antimicrobial, antitumor and insecticidal activities of
sucrose esters are important emerging areas of study, espe-
cially considering that such compounds are also likely to be
useful as food additives, drug carriers and/or emulsifiers.
Clearly, this is an evolving field of science where exciting
new discoveries are being made.
Funding
The authors thank the Program for Guangdong Pearl River
Introducing Innovative and Entrepreneurial Teams (grant
2017ZT07C571), the Program for Guangdong Yang Fan Introducing
Innovative and Entrepreneurial Teams (grant 2016YT03H132), the
Science and Technology Planning Project of Guangdong Province
(grant 2016A010105010) as well as the National Natural Science
Foundation of China (grant 21801094 and grant 31701525)
for funding.
ORCID
Yinglai Teng http://orcid.org/0000-0002-6260-8498
Scott G. Stewart http://orcid.org/0000-0002-7537-247X
Yao-Wen Hai http://orcid.org/0000-0003-3850-3239
Xuan Li http://orcid.org/0000-0001-8201-7751
Martin G. Banwell http://orcid.org/0000-0002-0582-475X
Ping Lan http://orcid.org/0000-0002-9285-3259
References
Abdel-Mageed, H. M., H. M. El-Laithy, L. G. Mahran, A. S. Fahmy, K.
M
ader, and S. A. Mohamed. 2012. Development of novel flexible
sugar ester vesicles as carrier systems for the antioxidant enzyme
catalase for wound healing applications. Process Biochemistry 47 (7):
115562. doi: 10.1016/j.procbio.2012.04.008.
Abouhilale, S., J. Greiner, and J. G. Riess. 1991. One-step preparation
of 6-perfluoroalkylalkanoates of trehalose and sucrose for biomedical
uses. Carbohydrate Research 212:5564. doi: 10.1016/0008-
6215(91)84045-G.
Akira, K., T. Taira, H. Hasegawa, C. Sakuma, and Y. Shinohara. 1998.
Studies on the stereoselective internal acyl migration of ketoprofen
glucuronides using
13
C labeling and nuclear magnetic resonance
spectroscopy. Drug Metabolism and Disposition: The Biological Fate
of Chemicals 26 (5):45764.
Anastas, P. T., and J. C. Warner. 1998. Green chemistry: Theory and
practice. New York: Oxford University Press.
Ariyaprakai, S. 2016. Formation and stability study of flavor oil emul-
sions stabilized by polyglycerol esters and by sucrose esters.
Assumption University of Thailand. Accessed December 31, 2019.
https://repository.au.edu/handle/6623004553/20047.
Ariyaprakai, S., X. Hu, and M. T. Tran. 2019. Spontaneous formation
of flavor oil emulsions by using sucrose esters and emulsion stability
study. Food Biophysics 14 (1):418. doi: 10.1007/s11483-018-9555-2.
Ariyaprakai, S., T. Limpachoti, and P. Pradipasena. 2013. Interfacial
and emulsifying properties of sucrose ester in coconut milk emul-
sions in comparison with Tween. Food Hydrocolloids 30 (1):35867.
doi: 10.1016/j.foodhyd.2012.06.003.
Ariyaprakai, S., and K. Tananuwong. 2015. Freeze-thaw stability of
edible oil-in-water emulsions stabilized by sucrose esters and
Tweens. Journal of Food Engineering 152:5764. doi: 10.1016/j.jfoo-
deng.2014.11.023.
Baczko, K., C. Nugier-Chauvin, J. Banoub, P. Thibault, and D.
Plusquellec. 1995. A new synthesis of 6-O-acylsucroses and of mixed
6,60-di-O-acylsucroses. Carbohydrate Research 269 (1):7988. doi: 10.
1016/0008-6215(94)00341-C.
Bai, J. 2009. Synthesis and bioactivity of sucrose selenite. MSc Thesis,
Central China Normal University, Wuhan, China. (in Chinese)
Bazin, H. G., T. Polat, and R. J. Linhardt. 1998. Synthesis of sucrose-
based surfactants through regioselective sulfonation of acylsucrose
and the nucleophilic opening of a sucrose cyclic sulfate.
16 Y. TENG ET AL.
Carbohydrate Research 309 (2):189205. doi: 10.1016/S0008-
6215(98)00121-9.
Belhaij, A., and O. Al-Mahdy. 2015. Foamability and foam stability of
several surfactants solutions: The role of screening and flooding.
Journal of Petroleum and Environmental Biotechnology 6 (4):
1000227.
Beuchat, L. R. 1980. Comparison of anti-vibrio activities of potassium
sorbate, sodium benzoate, and glycerol and sucrose esters of fatty
acids. Applied and Environmental Microbiology 39 (6):117882. doi:
10.1128/AEM.39.6.1178-1182.1980.
Bottle, S., and I. D. Jenkins. 1984. Improved synthesis of cord factor
analogues. Journal of the Chemical Society, Chemical
Communications (6):385. doi: 10.1039/C39840000385.
BRENDA. 2019a. BRENDA - Information on EC 2.4.1.9 - inulosucrase.
Accessed December 31, 2019. https://www.brenda-enzymes.org/
enzyme.php?ecno=2.4.1.9.
BRENDA. 2019b. BRENDA - Information on EC 3.1.1.3 - triacylgly-
cerol lipase. Accessed December 31, 2019. https://www.brenda-
enzymes.org/enzyme.php?ecno=3.1.1.3.
Buchanen, J. G., and D. A. Cummerson. 1972. 10,40:30,60-
Dianhydrosucrose. Carbohydrate Research 21 (2):293296. doi: 10.
1016/S0008-6215(00)82156-4.
Canadian Sugar Institute. 2019a. Global sugar trade (WTO). Accessed
December 31, 2019. https://sugar.ca/International-Trade/Global-
Sugar-Trade-(WTO).aspx.
Canadian Sugar Institute. 2019b. Sources of sugar. Accessed December
31, 2019. https://sugar.ca/Sugar-Basics/Sources-of-Sugar.aspx.
Cantor, S. M. 1939. Production of fatty acid esters from starch factory
byproducts. U.S. Patent No. RE21291E. Washington, DC: U.S.
Patent and Trademark Office.
Chamouleau, F., D. Coulon, M. Girardin, and M. Ghoul. 2001.
Influence of water activity and water content on sugar esters lipase-
catalyzed synthesis in organic media. Journal of Molecular Catalysis
B: Enzymatic 11 (46):949954. doi: 10.1016/S1381-1177(00)00166-1.
Chang, S. W., and J. F. Shaw. 2009. Biocatalysis for the production of
carbohydrate esters. New Biotechnology 26 (34):109116. doi: 10.
1016/j.nbt.2009.07.003.
Chauvin, C., K. Baczko, and D. Plusquellec. 1993. New highly regiose-
lective reactions of unprotected sucrose. Synthesis of 2-O-acylsucro-
ses and 2-O-(N-alkylcarbamoyl)sucroses. The Journal of Organic
Chemistry 58 (8):22912295. doi: 10.1021/jo00060a053.
Chauvin, C., and D. Plusquellec. 1991. A new chemoenzymatic synthe-
sis of 60-O-acylsucroses. Tetrahedron Letters 32 (29):34953498. doi:
10.1016/0040-4039(91)80815-N.
Chortyk, O. T. 2003. Chemically synthesized sugar esters for the con-
trol of soft-bodied arthropods. U.S. Patent No. 6,608,039.
Washington, DC: U.S. Patent and Trademark Office.
Chortyk, O. T., J. G. Pomonis, and A. W. Johnson. 1996. Syntheses
and characterizations of insecticidal sucrose esters. Journal of
Agricultural and Food Chemistry 44 (6):15511557. doi: 10.1021/
jf950615t.
Clode, D. M., D. McHale, J. B. Sheridan, G. G. Birch, and E. B.
Rathbone. 1985. Partial benzoylation of sucrose. Carbohydrate
Research 139 (1):141146. doi: 10.1016/0008-6215(85)90015-1.
Cruces, M. A., F. J. Plou, M. Ferrer, M. Bernab
e, and A. Ballesteros.
2001. Improved synthesis of sucrose fatty acid monoesters. Journal
of the American Oil ChemistsSociety 78 (5):541546. doi: 10.1007/
s11746-001-0300-5.
Dang, H. T., O. Obiri, and D. G. Hayes. 2005. Feed batch addition of
saccharide during saccharide-fatty acid esterification catalyzed by
immobilized lipase: Time course, water activity, and kinetic model.
Journal of the American Oil ChemistsSociety 82 (7):487493. doi:
10.1007/s11746-005-1098-x.
de Maria, P. D., J. M. S
anchez-Montero, J. V. Sinisterra, and A. R.
Alc
antara. 2006. Understanding Candida rugosa lipases: An over-
view. Biotechnology Advances 24 (2):180196. doi: 10.1016/j.biote-
chadv.2005.09.003.
Delange, R. J., and E. L. Smith. 1968. Substilisin Carlsberg. Journal of
Biological Chemistry 243 (9):21342142.
Deneyer, A., T. Ennaert, and B. F. Sels. 2018. Straightforward sustain-
ability assessment of sugar-derived molecules from first-generation
biomass. Current Opinion in Green and Sustainable Chemistry 10:
1120. doi: 10.1016/j.cogsc.2018.02.003.
Deshpande, P. S., T. D. Deshpande, R. D. Kulkarni, and P. P.
Mahulikar. 2013. Synthesis of sucrosecoconut fatty acids esters:
Reaction kinetics and rheological analysis. Industrial & Engineering
Chemistry Research 52 (43):1502415033. doi: 10.1021/ie401524g.
Devulapalle, K. S., A. G
omez de Segura, M. Ferrer, M. Alcalde, G.
Mooser, and F. J. Plou. 2004. Effect of carbohydrate fatty acid esters
on Streptococcus sobrinus and glucosyltransferase activity.
Carbohydrate Research 339 (6):10291034. doi: 10.1016/j.carres.2004.
01.007.
Du, L. H., and X. P. Luo. 2012. Lipase-catalyzed regioselective acylation
of sugar in microreactors. RSC Advances 2 (7):26632665. doi: 10.
1039/c2ra01112c.
Fameau, A. L., and A. Saint-Jalmes. 2017. Non-aqueous foams: Current
understanding on the formation and stability mechanisms. Advances
in Colloid and Interface Science 247:454464. doi: 10.1016/j.cis.2017.
02.007.
Fang, X., Z.-G. Zhuo, X.-K. Xu, J. Ye, H.-L. Li, Y.-H. Shen, and W. D.
Zhang. 2017. Cytotoxic isovaleryl sucrose esters from Ainsliaea yun-
nanensis: Reduction of mitochondrial membrane potential and
increase of reactive oxygen species levels in A549 cells. RSC
Advances 7 (34):2086520873. doi: 10.1039/C7RA01986F.
Farr
an, A., C. Cai, M. Sandoval, Y. Xu, J. Liu, M. J. Hern
aiz, and R. J.
Linhardt. 2015. Green solvents in carbohydrate chemistry: From raw
materials to fine chemicals. Chemical Reviews 115 (14):68116853.
doi: 10.1021/cr500719h.
Ferrer, M., M. A. Cruces, M. Bernab
e, A. Ballesteros, and F. J. Plou.
1999. Lipase-catalyzed regioselective acylation of sucrose in two-
solvent mixtures. Biotechnology and Bioengineering 65 (1):1016.
doi: 10.1002/(SICI)1097-0290(19991005)65:1<10::AID-BIT2>3.0.
CO;2-L.
Ferrer, M., F. Comelles, F. J. Plou, M. A. Cruces, G. Fuentes, J. L.
Parra, and A. Ballesteros. 2002a. Comparative surface activities of
di-and trisaccharide fatty acid esters. Langmuir 18 (3):667673. doi:
10.1021/la010727g.
Ferrer, M., F. J. Plou, G. Fuentes, M. A. Cruces, L. Andersen, O. Kirk,
M. Christensen, and A. Ballesteros. 2002b. Effect of the
Immobilization Method of Lipase from Thermomyces lanuginosus on
Sucrose Acylation. Biocatalysis and Biotransformation 20 (1):6371.
doi: 10.1080/10242420210153.
Ferrer, M., J. Soliveri, F. J. Plou, N. L
opez-Cort
es, D. Reyes-Duarte, M.
Christensen, J. L. Copa-Pati~
no, and A. Ballesteros. 2005. Synthesis of
sugar esters in solvent mixtures by lipases from Thermomyces lanu-
ginosus and Candida antarctica B, and their antimicrobial properties.
Enzyme and Microbial Technology 36 (4):391398. doi: 10.1016/j.
enzmictec.2004.02.009.
Feuge, R. O., H. J. Zeringue, T. J. Weiss, and M. Brown. 1970.
Preparation of sucrose esters by interesterification. Journal of the
American Oil ChemistsSociety 47 (2):5660. doi: 10.1007/
BF02541458.
Filho, D. G., A. G. Silva, and C. Z. Guidini. 2019. Lipases: Sources,
immobilization methods, and industrial applications. Applied
Microbiology and Biotechnology 103 (18):73997423. doi: 10.1007/
s00253-019-10027-6.
Fischer, F., M. Happe, J. Emery, A. Fornage, and R. Sch
utz. 2013.
Enzymatic synthesis of 6- and 60-O-linoleyl-a-D-maltose: From solv-
ent-free to binary ionic liquid reaction media. Journal of Molecular
Catalysis B: Enzymatic 90:98106. doi: 10.1016/j.molcatb.2013.01.
019.
Food Standards Agency. 2018. EU approved additives and E numbers.
Accessed December 31, 2019. https://www.food.gov.uk/business-
guidance/eu-approved-additives-and-e-numbers.
Franssen, M. C. R., P. Steunenberg, E. L. Scott, H. Zuilhof, and
J. P. M. Sanders. 2013. Immobilised enzymes in biorenewables pro-
duction. Chemical Society Reviews 42 (15):64916533. doi: 10.1039/
c3cs00004d.
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 17
Garofalakis, G., and B. S. Murray. 2001. Dilatational rheology and
foaming properties of sucrose monoesters in the presence of b-lacto-
globulin. Colloids and Surfaces. B, Biointerfaces 21 (13):317. doi:
10.1016/S0927-7765(01)00179-5.
Goderis, H. L., G. Ampe, M. P. Feyten, B. L. Fouw
e, W. M. Guffens,
S. M. Van Cauwenbergh, and P. P. Tobback. 1987. Lipase-catalyzed
ester exchange reactions in organic media with controlled humidity.
Biotechnology and Bioengineering 30 (2):258266. doi: 10.1002/bit.
260300216.
Griffin, W. C. 1949. Classification of surface-active agents by HLB.
Journal of the Society of Cosmetic Chemists 1:311326.
Griffin, W. C. 1954. Calculation of HLB values of non-ionic surfac-
tants. Journal of the Society of Cosmetic Chemists 5 (4):249256.
Guan, Y., H. Chen, and Q. Zhong. 2019. Nanoencapsulation of caffeic
acid phenethyl ester in sucrose fatty acid esters to improve activities
against cancer cells. Journal of Food Engineering 246:125133. doi:
10.1016/j.jfoodeng.2018.11.008.
Gumel, A. M., M. S. M. Annuar, T. Heidelberg, and Y. Chisti. 2011.
Lipase mediated synthesis of sugar fatty acid esters. Process
Biochemistry 46 (11):20792090. doi: 10.1016/j.procbio.2011.07.021.
Gupta, R. K., K. James, and F. J. Smith. 1983. Sucrose esters and
sucrose ester/glyceride blends as emulsifiers. Journal of the American
Oil ChemistsSociety 60 (4):862869. doi: 10.1007/BF02787451.
Guthrie, R. D., I. D. Jenkins, and R. Yamasaki. 1980. Epoxidation with
triphenylphosphine and diethyl azodicarboxylate: Epoxide derivatives
of sucrose. Carbohydrate Research 85 (2):C5C6. doi: 10.1016/
S0008-6215(00)84683-2.
Guti
errez, M. F., J. L. Rivera, A. Suaza, and A. Orjuela. 2018. Kinetics
of the transesterification of methyl palmitate and sucrose using sur-
factants. Chemical Engineering Journal and the Biochemical
Engineering Journal 347:877888. doi: 10.1016/j.cej.2018.04.085.
Habulin, M., S.
Sabeder, and
Z. Knez. 2008. Enzymatic synthesis of
sugar fatty acid esters in organic solvent and in supercritical carbon
dioxide and their antimicrobial activity. Journal of Supercritical
Fluids 45 (3):338345. doi: 10.1016/j.supflu.2008.01.002.
Haines, A. H. 1976. Relative reactivities of hydroxyl groups in carbohy-
drates. In Advances in Carbohydrate Chemistry and Biochemistry, ed.
R. S. Tipson, and D. Horton, vol. 33, 11109. New York, USA:
Academic Press.
Herzfeld, A. 1880. Acetylirung einiger Kohlenhydrate nach dem
Liebermannschen Verfahren. Berichte d. D. chem. Gesellschaft (mer-
gered into. Berichte Der Deutschen Chemischen Gesellschaft 13 (1):
265268. doi: 10.1002/cber.18800130175.
Hill, K., and O. Rhode. 1999. Sugar-based surfactants for consumer
products and technical applications. Lipid - Fett 101 (1):2533. doi:
10.1002/(SICI)1521-4133(19991)101:1<25::AID-LIPI25>3.0.CO;2-N.
Huang, D., X. Jiang, H. Zhu, X. Fu, K. Zhong, and W. Gao. 2010a.
Improved synthesis of sucrose fatty acid monoesters under ultra-
sonic irradiation. Ultrasonics Sonochemistry 17 (2):352355. doi: 10.
1016/j.ultsonch.2009.08.009.
Huang, G., T. Wang, P. Yao, and Y. Wei. 2010b. Studies on enzymatic
synthesis of 6-O-lauroylsucrose in ionic liquids. Journal of Guangxi
University 35 (3):455461. (in Chinese)
Human Metabolome Database. 2019. Showing metabocard for Sucrose
(HMDB0000258). Accessed December 31, 2019. http://www.hmdb.
ca/metabolites/HMDB0000258.
Husband, F. A., D. B. Sarney, M. J. Barnard, and P. J. Wilde. 1998.
Comparison of foaming and interfacial properties of pure sucrose
monolaurates, dilaurate and commercial preparations. Food
Hydrocolloids 12 (2):237244. doi: 10.1016/S0268-005X(98)00036-8.
Hutzler, S., D. L
osch, E. Carey, D. Weaire, M. Hloucha, and C.
Stubenrauch. 2011. Evaluation of a steady-state test of foam stability.
Philosophical Magazine 91 (4):537552. doi: 10.1080/14786435.2010.
526646.
Ikekawa, T., M. Umeji, S. Yanoma, K. Yoshimoto, and Y. Nishikawa.
1979. Chemical and biochemical studies on carbohydrate esters.
VIII. Antitumor activity of sucrose fatty acid esters. Chemical &
Pharmaceutical Bulletin 27 (9):20162020. doi: 10.1248/cpb.27.2016.
Itoh, T. 2017. Ionic liquids as tool to improve enzymatic organic syn-
thesis. Chemical Reviews 117 (15):1056710607. doi: 10.1021/acs.
chemrev.7b00158.
Iwami, Y., C. F. Schachtele, and T. Yamada. 1995. Effect of sucrose
monolaurate on acid production, levels of glycolytic intermediates,
and enzyme activities of streptococcus mutans NCTC 10449. Journal
of Dental Research 74 (9):16131617. doi: 10.1177/
00220345950740091801.
Jackson, D. M., O. T. Chortyk, M. G. Stephenson, A. W. Johnson,
C. D. Harlow, A. M. Simmons, and V. A. Sisson. 1998. Potential of
Nicotiana species for production of sugar esters. Tobacco Science 42:
19.
Jiang, J., W. Jing, Y. L. Xiong, and Y. Liu. 2019. Interfacial competitive
adsorption of different amphipathicity emulsifiers and milk protein
affect fat crystallization, physical properties, and morphology of fro-
zen aerated emulsion. Food Hydrocolloids 87:670678. doi: 10.1016/j.
foodhyd.2018.09.005.
Johnson, K. E. 2007. What is an ionic liquid? Electrochemical Society
Interface 16 (1):3841.
Kato, A., K. Ando, G. A. Jamura, and K. Arima. 1971. Effects of some
fatty acid esters on the viability and transplantability of Ehrlich asci-
tes tumor cells. Cancer Resources 31:501504.
Kea, S., Walker, C. E., and Kline, E. 1987. Sugar esters and an
improved anhydrous method of manufacture. U.S. Patent No.
4,683,299. Washington, DC: U.S. Patent and Trademark Office.
Knez,
Z. 2018. Enzymatic reactions in subcritical and supercritical flu-
ids. Journal of Supercritical Fluids 134:133140. doi: 10.1016/j.supflu.
2017.11.023.
Kowalski, M., P. Cmoch, and S. Jarosz. 2014. Novel approach to select-
ively functionalized derivatives of sucrose. Synlett 25 (05):641644.
doi: 10.1055/s-0033-1340180.
Le Maire, M., P. Champeil, and J. V. Møller. 2000. Interaction of mem-
brane proteins and lipids with solubilizing detergents. Biochimica et
Biophysica Acta, Biomembranes 1508 (12):86111. doi: 10.1016/
S0304-4157(00)00010-1.
Li, G., Y. Cai, Z. Hao, and X. Liao. 2011a. Synthesis of sucrose acetate
using a solvent-stable serine protease from Serratia sp. SYBC H.
Engineering in Life Sciences 11 (6):615619. doi: 10.1002/elsc.
201100038.
Li, G., Y. Cai, X. Liao, and J. Yin. 2011b. Enzymatic synthesis of
sucrose octaacetate using a novel alkaline protease. Biotechnology
Letters 33 (3):607610. doi: 10.1007/s10529-010-0468-5.
Li, X., Y. W. Hai, D. Ma, J. Chen, M. G. Banwell, and P. Lan. 2019.
Fatty acid ester surfactants derived from raffinose: Synthesis, charac-
terization and structure-property profiles. Journal of Colloid and
Interface Science 556:616627. doi: 10.1016/j.jcis.2019.08.070.
Liang, M. Y., M. G. Banwell, Y. Wang, and P. Lan. 2018a. Effect of var-
iations in the fatty acid residue of lactose monoesters on their emul-
sifying properties and biological activities. Journal of Agricultural
and Food Chemistry 66 (47):1259412603. doi: 10.1021/acs.jafc.
8b05794.
Liang, M. Y., Y. Chen, M. G. Banwell, Y. Wang, and P. Lan. 2018b.
Enzymatic preparation of a homologous series of long-chain 6-O-
acylglucose esters and their evaluation as emulsifiers. Journal of
Agricultural and Food Chemistry 66 (15):39493956. doi: 10.1021/
acs.jafc.8b00913.
Lie, A., A. S. Meyer, and L. H. Pedersen. 2014. Appearance and distri-
bution of regioisomers in metallo- and serine-protease-catalysed
acylation of sucrose in N,N-dimethylformamide. Journal of
Molecular Catalysis B: Enzymatic 106:2631. doi: 10.1016/j.molcatb.
2014.04.017.
Lin, C., H. Cui, X. Wang, H. Wang, S. Xia, K. Hayat, S. Hussain,
M. U. Tahir, and X. Zhang. 2020. Regulating water binding capacity
and improving porous carbohydrate matrixs humectant and mois-
ture proof functions by mixture of sucrose ester and Polygonatum
sibiricum polysaccharide. International Journal of Biological
Macromolecules 147:667674. doi: 10.1016/j.ijbiomac.2020.01.101.
Liu, Q., M. H. A. Janssen, F. van Rantwijk, and R. A. Sheldon. 2005.
Room-temperature ionic liquids that dissolve carbohydrates in high
concentrations. Green Chemistry 7 (1):3942. doi: 10.1039/b412848f.
18 Y. TENG ET AL.
Liu, T.-X., P. A. Stansly, and O. T. Chortyk. 1996. Insecticidal activity
of natural and synthetic sugar esters against Bemisia argentifolii
(homopetera: Aleyrodidae). Journal of Economic Entomology 89 (5):
12331239. doi: 10.1093/jee/89.5.1233.
Ma, Y.-R., M. G. Banwell, R. Yan, and P. Lan. 2018. Comparative study
of the emulsifying properties of a homologous series of long-chain
60-O-acylmaltose esters. Journal of Agricultural and Food Chemistry
66 (33):88328840. doi: 10.1021/acs.jafc.8b02391.
MarketsandMarkets. 2019a. Global Sucrose Esters Market Segment
Outlook, Market Assessment, Competition Scenario, Trends and
Forecast 2019-2028. Accessed December 31, 2019. https://market.us/
report/sucrose-esters-market.
MarketsandMarkets. 2019b. Sucrose esters market worth 74.6 million
USD by 2020. Accessed December 31, 2019.https://www.marketsand-
markets.com/PressReleases/sucrose-esters.asp.
Marshall, D. L., and L. B. Bullerman. 1986. Antimicrobial activity of
sucrose fatty acid ester emulsifiers. Journal of Food Science 51 (2):
468470. doi: 10.1111/j.1365-2621.1986.tb11157.x.
Martinelle, M., and K. Hult. 1995. Kinetics of acyl transfer reactions in
organic media catalysed by Candida antarctica lipase B. Biochimica
et Biophysica Acta (BBA) - Protein Structure and Molecular
Enzymology 1251 (2):1917. doi: 10.1016/0167-4838(95)00096-D.
McClements, D. J. 2015. Emulsion Properties. In D. J. McClements
(Ed.), Food emulsions: principles, practices, and techniques (820).
Boca Raton: CRC Press.
McKenzie, C. L., and G. J. Puterka. 2004. Effect of sucrose octanoate
on survival of nymphal and adult Diaphorina citri (Homoptera:
Psyllidae). Journal of Economic Entomology 97 (3):9705.2.0.co;2]
[15279280]
Mitsubishi Chemicals. 2019. Sucrose fatty acid ester Ryoto
TM
Sugar
Ester. Cemerlang, M. J. 2019. Accessed December 31, 2019. https://
www.m-chemical.co.jp/en/products/departments/group/mfc/product/
1201443_7739.html.
Moh, M. H., T. S. Tang, and G. H. Tan. 2000. Improved separation of
sucrose ester isomers using gradient high performance liquid chro-
matography with evaporative light scattering detection. Food
Chemistry 69 (1):10510. doi: 10.1016/S0308-8146(99)00226-5.
Molinier, V., J. Fitremann, A. Bouchu, and Y. Queneau. 2004. Sucrose
esterification under Mitsunobu conditions: Evidence for the forma-
tion of 6-O-acyl-30,60-anhydrosucrose besides mono and diesters of
fatty acids. Tetrahedron: Asymmetry 15 (11):175362. doi: 10.1016/j.
tetasy.2004.04.021.
Molinier, V., K. Wisniewski, A. Bouchu, J. Fitremann, and Y. Queneau.
2003. Transesterification of sucrose in organic medium: Study of
acyl group migrations. Journal of Carbohydrate Chemistry 22 (78):
65769. doi: 10.1081/CAR-120026466.
Monk, J. D., L. R. Beuchat, and A. K. Hathcox. 1996. Inhibitory effects
of sucrose monolaurate, alone and in combination with organic
acids, on Listeria monocytogenes and Staphylococcus aureus.Journal
of Applied Bacteriology 81 (1):718. doi: 10.1111/j.1365-2672.1996.
tb03276.x.
Nakayama, M., D. Tomiyama, K. Ikeda, M. Katsuki, A. Nonaka, and T.
Miyamoto. 2015. Antibacterial effects of monoglycerol fatty acid
esters and sucrose fatty acid esters on bacillus spp. Food Science and
Technology Research 21 (3):4317. doi: 10.3136/fstr.21.431.
Neta, N. A. S., J. C. S. dos Santos, S. O. Sancho, S. Rodrigues, L. R. B.
Gonc¸alves, L. R. Rodrigues, and J. A. Teixeira. 2012. Enzymatic syn-
thesis of sugar esters and their potential as surface-active stabilizers
of coconut milk emulsions. Food Hydrocolloids 27 (2):32431. doi:
10.1016/j.foodhyd.2011.10.009.
Neta, N. S., J. A. Teixeira, and L. R. Rodrigues. 2015. Sugar ester sur-
factants: Enzymatic synthesis and applications in food industry.
Critical Reviews in Food Science and Nutrition 55 (5):595610. doi:
10.1080/10408398.2012.667461.
Nishikawa, Y., M. Okabe, K. Yoshimoto, G. Kurono, and F. Fukuoka.
1976a. Chemical and biochemical studies on carbohydrate esters. II.
Antitumor activity of saturated fatty acids and their ester derivatives
against Ehrlich ascites carcinoma. Chemical & Pharmaceutical
Bulletin 24 (3):38793. doi: 10.1248/cpb.24.387.
Nishikawa, Y., K. Yoshimoto, M. Okabe, and F. Fukuoka. 1976b.
Chemical and biochemical studies on carbohydrate esters. III.
Antitumor activity of unsaturated fatty acids and their ester deriva-
tives against Ehrlich ascites carcinoma. Chemical & Pharmaceutical
Bulletin 24 (4):75662. doi: 10.1248/cpb.24.756.
Nishikawa, Y., K. Yoshimoto, M. Okada, T. Ikekawa, N. Abiko, and F.
Fukuoka. 1977a. Chemical and biochemical studies on carbohydrate
esters. V. Anti Ehrlich ascites tumor effect and chromatographic
behaviors of fatty acyl monoesters of sucrose and trehalose.
Chemical & Pharmaceutical Bulletin 25 (7):171724. doi: 10.1248/
cpb.25.1717.
Nishikawa, Y., K. Yoshimoto, T. Manabe, and T. Ikekawa. 1977b.
Chemical and biochemical studies on carbohydrate esters. VI.
Further examinations on antitumor activities of stearoyl esters of
sucrose. Chemical & Pharmaceutical Bulletin 25 (9):237884. doi: 10.
1248/cpb.25.2378.
Ortiz-Soto, M. E., C. Possiel, J. G
orl, A. Vogel, R. Schmiedel, and J.
Seibel. 2017. Impaired coordination of nucleophile and increased
hydrophobicity in the þ1 subsite shift levansucrase activity towards
transfructosylation. Glycobiology 27 (8):75565. doi: 10.1093/glycob/
cwx050.
Otomo, N., 2009. Basic properties of sucrose fatty acid esters and their
applications. Biobased surfactants and detergents: Synthesis, proper-
ties, and applications, ed. D. G. Hayes, D. Kitamoto, D. K. Y.
Solaiman and R. D. Ashby, 27598. Urbana, Illinois: AOCS Press.
Park, K.-M., S. J. Lee, H. Yu, J.-Y. Park, H.-S. Jung, K. Kim, C. J. Lee,
and P.-S. Chang. 2018. Hydrophilic and lipophilic characteristics of
non-fatty acid moieties: Significant factors affecting antibacterial
activity of lauric acid esters. Food Science and Biotechnology 27 (2):
4019. doi: 10.1007/s10068-018-0353-x.
Parker, K. J., R. A. Khan, and W. I. C. York. 1973. Process for the pro-
duction of surface active agents comprising sucrose esters. Great
Britain Patent No. 1,399,053A. London: Intellectual Property Office
of the United Kingdom.
Parker, K. J., Khan, R. A., and Mufti, K. S. 1976. Process of making
sucrose esters. U.S. Patent No. 3,996,206. Washington, DC: U.S.
Patent and Trademark Office.
Patel, A. R. 2017. Stable arrestednon-aqueous edible foams based on
food emulsifiers. Food & Function 8 (6):211520. doi: 10.1039/
c7fo00187h.
Patil, D. R., D. G. Rethwisch, and J. S. Dordick. 1991. Enzymatic syn-
thesis of a sucrose-containing linear polyester in nearly anhydrous
organic media. Biotechnology and Bioengineering 37 (7):63946. doi:
10.1002/bit.260370706.
Pearce, K. N., and J. E. Kinsella. 1978. Emulsifying properties of pro-
teins: Evaluation of a turbidimetric technique. Journal of
Agricultural and Food Chemistry 26 (3):71623. doi: 10.1021/
jf60217a041.
Pedersen, N. R., P. J. Halling, L. H. Pedersen, R. Wimmer, R.
Matthiesen, and O. R. Veltman. 2002. Efficient transesterification of
sucrose catalysed by the metalloprotease thermolysin in dimethyl-
sulfoxide. FEBS Letters 519 (13):1814. doi: 10.1016/S0014-
5793(02)02753-9.
Pedersen, N. R., R. Wimmer, R. Matthiesen, L. H. Pedersen, and A.
Gessesse. 2003. Synthesis of sucrose laurate using a new alkaline
protease. Tetrahedron: Asymmetry 14 (6):66773. doi: 10.1016/
S0957-4166(03)00086-7.
Petkova, N., D. Vassilev, R. Grudeva, Y. Tumbarski, I. Vasileva, M.
Koleva, and P. Denev. 2017. Greensynthesis of sucrose octaacetate
and characterization of its physicochemical properties and anti-
microbial activity. Chemical and Biochemical Engineering Quarterly
31 (4):395402. doi: 10.15255/CABEQ.2017.1117.
Petrova, K. T., M. T. Barros, R. C. Calhelha, M. Sokovi
c, and I. C.
Ferreira. 2018. Antimicrobial and cytotoxic activities of short carbon
chain unsaturated sucrose esters. Medicinal Chemistry Research 27
(3):9808. doi: 10.1007/s00044-017-2121-5.
Plou, F. J., M. A. Cruces, M. Ferrer, G. Fuentes, E. Pastor, M. Bernab
e,
M. Christensen, F. Comelles, J. L. Parra, and A. Ballesteros. 2002.
Enzymatic acylation of di- and trisaccharides with fatty acids:
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 19
Choosing the appropriate enzyme, support and solvent. Journal of
Biotechnology 96 (1):5566. doi: 10.1016/S0168-1656(02)00037-8.
Polat, T., H. G. Bazin, and R. J. Linhardt. 1997. Enzyme catalyzed
regioselective synthesis of sucrose fatty acid ester surfactants. Journal
of Carbohydrate Chemistry 16 (9):131925. doi: 10.1080/
07328309708005752.
Polat, T., and R. J. Linhardt. 2001. Syntheses and applications of
sucrose-based esters. Journal of Surfactants and Detergents 4 (4):
41521. doi: 10.1007/s11743-001-0196-y.
Possiel, C., A. B
auerle, and J. Seibel. 2017. A chemoenzymatic route to
a class of sucrose esters. European Journal of Organic Chemistry
2017 (42):63357. doi: 10.1002/ejoc.201701313.
Potier, P., V. Maccario, M. B. Giudicelli, Y. Queneau, and O. Dangles.
1999. Gallic esters of sucrose as a new class of antioxidants.
Tetrahedron Letters 40 (17):338790. doi: 10.1016/S0040-
4039(99)00496-7.
Puterka, G. J., W. Farone, T. Palmer, and A. Barrington. 2003.
Structure-function relationships affecting the insecticidal and mitici-
dal activity of sugar esters. Journal of Economic Entomology 96 (3):
63644. doi: 10.1093/jee/96.3.636.
Queneau, Y., S. Chambert, C. Besset, and R. Cheaib. 2008. Recent pro-
gress in the synthesis of carbohydrate-based amphiphilic materials:
The examples of sucrose and isomaltulose. Carbohydrate Research
343 (12):19992009. doi: 10.1016/j.carres.2008.02.008.
Queneau, Y., J. Fitremann, and S. Trombotto. 2004. The chemistry of
unprotected sucrose: The selectivity issue. Comptes Rendus Chimie 7
(2):17788. doi: 10.1016/j.crci.2003.10.014.
Rawlings, N. D., A. J. Barrett, P. D. Thomas, X. Huang, A. Bateman,
and R. D. Finn. 2018. The MEROPS database of proteolytic
enzymes, their substrates and inhibitors in 2017 and a comparison
with peptidases in the PANTHER database. Nucleic Acids Research
46 (D1):D624D632. doi: 10.1093/nar/gkx1134.
Ren, K., and B. P. Lamsal. 2017. Synthesis of some glucose-fatty acid
esters by lipase from Candida antarctica and their emulsion func-
tions. Food Chemistry 214:55663. doi: 10.1016/j.foodchem.2016.07.
031.
Reyes-Duarte, D., N. L
opez-Cort
es, M. Ferrer, F. J. Plou, and A.
Ballesteros. 2005. Parameters affecting productivity in the lipase-cat-
alysed synthesis of sucrose palmitate. Biocatalysis and
Biotransformation 23 (1):1927. doi: 10.1080/10242420500071763.
Riva, S., J. Chopineau, A. P. G. Kieboom, and A. M. Klibanov. 1988.
Protease-catalyzed regioselective esterification of sugars and related
compounds in anhydrous dimethylformamide. Journal of the
American Chemical Society 110 (2):5849. doi: 10.1021/ja00210a045.
Roshdy, I., Horst, C., and Herbert, S. 1967. Process for the production
of sucrose esters of fatty acids. U.S. Patent No. 3,347,848.
Washington, DC: U.S. Patent and Trademark Office.
Royal Society of Chemistry. 2019. ChemSpider. Accessed December 31,
2019. http://www.chemspider.com.
Salihu, A., and M. Z. Alam. 2015. Solvent tolerant lipases: A review.
Process Biochemistry 50 (1):8696. doi: 10.1016/j.procbio.2014.10.
019.
Satyawali, Y., K. Vanbroekhoven, and W. Dejonghe. 2017. Process
intensification: The future for enzymatic processes? Biochemical
Engineering Journal 121:196223. doi: 10.1016/j.bej.2017.01.016.
Schomburg, I., L. Jeske, M. Ulbrich, S. Placzek, A. Chang, and D.
Schomburg. 2017. The BRENDA enzyme information system-From
a database to an expert system. Journal of Biotechnology 261:
194206. doi: 10.1016/j.jbiotec.2017.04.020.
Severson, R. F., O. T. Chortyk, M. G. Stephenson, D. H. Akey, J. W.
Neal, G. W. Pittarelli, D. M. Jackson, and V. A. Sisson. 1994.
Characterization of Natural Pesticide from Nicotiana gossei in
Bioregulators for Crop Protection and Pest Control.ACS Symposium
Series, Vol. 557, Chapter 9, 10921. Washington D.C., US: ACS
Publications.
Shao, S.-Y., Y.-G. Shi, Y. Wu, L.-Q. Bian, Y.-J. Zhu, X.-Y. Huang, Y.
Pan, L.-Y. Zeng, and R.-R. Zhang. 2018. Lipase-catalyzed synthesis
of sucrose monolaurate and its antibacterial property and mode of
action against four pathogenic bacteria. Molecules 23 (5):111835.
doi: 10.3390/molecules23051118.
Shearer, A. E. H., C. P. Dunne, A. Sikes, and D. G. Hoover. 2000.
Bacterial spore inhibition and inactivation in foods by pressure,
chemical preservatives, and mild heat. Journal of Food Protection 63
(11):150310. doi: 10.4315/0362-028X-63.11.1503.
Shi, Y-g, J-r Li, and Y.-H. Chu. 2011. Enzyme-catalyzed regioselective
synthesis of sucrose-based esters. Journal of Chemical Technology &
Biotechnology 86 (12):145768. doi: 10.1002/jctb.2711.
Singh, S. K. 2019. Solubility of lignin and chitin in ionic liquids and
their biomedical applications. International Journal of Biological
Macromolecules 132:26577. doi: 10.1016/j.ijbiomac.2019.03.182.
Sisterna. 2019a. DK ester sucrose esters of fatty acids. Accessed
August 22, 2019. http://www.mjc.biz.id/dk-ester.html.
Sisterna. 2019b. Other sucrose esters. Accessed December 31, 2019.
https://www.sisterna.com/about-sisterna/our-sucrose-esters.
Soultani, S., S. Ognier, J.-M. Engasser, and M. Ghoul. 2003.
Comparative study of some surface active properties of fructose
esters and commercial sucrose esters. Colloids and Surfaces A:
Physicochemical and Engineering Aspects 227 (13):3544. doi: 10.
1016/S0927-7757(03)00360-1.
Smith, C. E., and Tuschhoff, J. V. 1957. Acylation of hydroxy com-
pounds with vinyl esters. U.S. Patent No. 2,928,828. Washington,
DC: U.S. Patent and Trademark Office.
Soares, A., S. de, P. E. D. Augusto, B. R. J
unior, C. L. de, C. A.
Nogueira,
E. N. R. Vieira, F. A. R. de Barros, P. C. Stringheta, and
A. M. Ramos. 2019. Ultrasound assisted enzymatic hydrolysis of
sucrose catalyzed by invertase: Investigation on substrate, enzyme
and kinetics parameters. Lwt 107:16470. doi: 10.1016/j.lwt.2019.02.
083.
Song, Z., S. Li, X. Chen, L. Liu, and Z. Song. 2006. Synthesis of insecti-
cidal sucrose esters. Forestry Studies in China 8 (3):269. doi: 10.
1007/s11632-006-0019-2.
Sugimoto, K., H. Tanaka, R. Moriyama, Y. Nagai, A. Ogawa, and S.
Makino. 1998. A loss of the antimicrobial activity of sucrose mono-
esters of fatty acids as caused by esterase released from the germi-
nated spores and vegetative cells of Bacillus cereus.Biocontrol
Science 3 (1):1721. doi: 10.4265/bio.3.17.
Szuts, A., M. Budai-Szucs, I. Eros, N. Otomo, and P. Szab
o-R
ev
esz.
2010. Study of gel-forming properties of sucrose esters for thermo-
sensitive drug delivery systems. International Journal of
Pharmaceutics 383 (12):1327. doi: 10.1016/j.ijpharm.2009.09.013.
Teranishi, K. 2002. Direct regioselective 2-O-(p-toluenesulfonylation) of
sucrose. Carbohydrate Research 337 (7):6139. doi: 10.1016/S0008-
6215(02)00039-3.
Th
evenet, S., G. Descotes, A. Bouchu, and Y. Queneau. 1997.
Hydrophobic effect driven esterification of sucrose in aqueous
medium. Journal of Carbohydrate Chemistry 16 (45):6916. doi: 10.
1080/07328309708007347.
Th
evenet, S., A. Wernicke, S. Belniak, G. Descotes, A. Bouchu, and Y.
Queneau. 1999. Esterification of unprotected sucrose with acid
chlorides in aqueous medium: Kinetic reactivity versus acyl- or alky-
loxycarbonyl-group migrations. Carbohydrate Research 318 (14):
5266. doi: 10.1016/S0008-6215(99)00079-8.
Thomas, L. V., E. A. Davies, J. Delves-Broughton, and J. W.
Wimpenny. 1998. Synergist effect of sucrose fatty acid esters on
nisin inhibition of Gram-positive bacteria. Journal of Applied
Microbiology 85 (6):101322. doi: 10.1111/j.1365-2672.1998.tb05266.
x.
Tual, A., E. Bourles, P. Barey, A. Houdoux, M. Desprairies, and J.-L.
Courthaudon. 2006. Effect of surfactant sucrose ester on physical
properties of dairy whipped emulsions in relation to those of O/W
interfacial layers. Journal of Colloid and Interface Science 295 (2):
495503. doi: 10.1016/j.jcis.2005.09.010.
Tucker, N. B., and Martin, J. B. 1958. Method for preparing fatty esters
of non-reducing oligosaccharides in the presence of an amide. U.S.
Patent No. 2,831,854. Washington, DC: U.S. Patent and Trademark
Office.
Vassilev, D., N. Petkova, M. Koleva, and P. Denev. 2016. Ultrasound-
assisted synthesis of sucrose and fructooligosaccharides esters as bio-
plasticizers. Journal of Renewable Materials 4 (1):2430. doi: 10.
7569/JRM.2015.634125.
20 Y. TENG ET AL.
Verkempinck, S. H. E., L. Salvia-Trujillo, L. G. Moens, L. Charleer,
A. M. van Loey, M. E. Hendrickx, and T. Grauwet. 2018. Emulsion
stability during gastrointestinal conditions effects lipid digestion kin-
etics. Food Chemistry 246:17991. doi: 10.1016/j.foodchem.2017.11.
001.
Vlahov, I. R., P. I. Vlahova, and R. J. Linhardt. 1997. Regioselective
synthesis of sucrose monoesters as surfactants. Journal of
Carbohydrate Chemistry 16 (1):110. doi: 10.1080/
07328309708006506.
Wagh, A., S. Shen, F. A. Shen, C. D. Miller, and M. K. Walsh. 2012.
Effect of lactose monolaurate on pathogenic and nonpathogenic bac-
teria. Applied and Environmental Microbiology 78 (9):34658. doi:
10.1128/AEM.07701-11.
Walsh, M. K., R. A. Bombyk, A. Wagh, A. Bingham, and L. M.
Berreau. 2009. Synthesis of lactose monolaurate as influenced by
various lipases and solvents. Journal of Molecular Catalysis B:
Enzymatic 60 (34):1717. doi: 10.1016/j.molcatb.2009.05.003.
Walstra, P. 2002. Physical chemistry of foods. New York: CRC Press.
Wang, X., S. Miao, P. Wang, and S. Zhang. 2012. Highly efficient syn-
thesis of sucrose monolaurate by alkaline protease Protex 6L.
Bioresource Technology 109:712. doi: 10.1016/j.biortech.2012.01.035.
Wang, Q., S. Zhang, and J. Yang. 2007. Regioselective formation of 6-
O-acylsucroses and 6,3-di-O-acylsucroses via the stannylene acetal
method. Carbohydrate Research 342 (17):265763. doi: 10.1016/j.
carres.2007.08.014.
Wei, T., W. Jia, X. Yu, and D. Mao. 2016. Enhancement of enzymatic
synthesis of sucrose 6-acetate with Aspergillus oryzae fructosyltrans-
ferase using ionic liquid as a cosolvent. Journal of Molecular
Catalysis B: Enzymatic 123:1006. doi: 10.1016/j.molcatb.2015.11.
016.
Wei, T., X. Yu, Y. Wang, Y. Zhu, C. Du, C. Jia, and D. Mao. 2014.
Purification and evaluation of the enzymatic properties of a novel
fructosyltransferase from Aspergillus oryzae: A potential biocatalyst
for the synthesis of sucrose 6-acetate. Biotechnology Letters 36 (5):
101520. doi: 10.1007/s10529-014-1457-x.
Weiss, T. J., M. Brown, H. J. Zeringue, Jr., and R. O. Feuge. 1971.
Quantitative estimation of sucrose esters of plamitic acid. Journal of
the American Oil ChemistsSociety 48 (4):1458. doi: 10.1007/
BF02544937.
Welton, T. 1999. Room-temperature ionic liquids. Solvents for synthe-
sis and catalysis. Chemical Reviews 99 (8):207184. doi: 10.1021/
cr980032t.
Wuts, P. G. M., and T. W. Greene. 2006. Protection for the hydroxyl
group, including 1,2- and 1,3-diols. In P. G. M. Wuts and T. W.
Greene (Ed.), Greenes protective groups in organic synthesis, 4 ed.(
16366. Hoboken, New Jersey: John Wiley & Sons. 10.1002/
0471220574.ch2.
Xiao, D., R. Ye, P. M. Davidson, D. G. Hayes, D. A. Golden, and Q.
Zhong. 2011. Sucrose monolaurate improves the efficacy of sodium
hypochlorite against Escherichia coli O157:H7 on spinach.
International Journal of Food Microbiology 145 (1):648. doi: 10.
1016/j.ijfoodmicro.2010.11.029.
Yamashita, S., K. Matsumiya, Y. Kogo, K. Takamatsu, and Y.
Matsumura. 2020. Emulsifiers efficiently prevent hardening of pan-
cakes under refrigerated conditions via inclusion complexes with
starch molecules. Food Hydrocolloids 100:105432. doi: 10.1016/j.
foodhyd.2019.105432.
Yang, C.-M., L. O. Luedecke, B. G. Swanson, and P. M. Davidson.
2003. Inhibition of microorganisms in salad dressing by sucrose and
methylglucose fatty acid monoesters. Journal of Food Processing and
Preservation 27 (4):28598. doi: 10.1111/j.1745-4549.2003.tb00518.x.
Yang, Z., and Z.-L. Huang. 2012. Enzymatic synthesis of sugar fatty
acid esters in ionic liquids. Catalysis Science & Technology 2 (9):
176775. doi: 10.1039/c2cy20109g.
Ye, R., and D. G. Hayes. 2014. Recent progress for lipase-catalysed syn-
thesis of sugar fatty acid esters. Journal of Oil Palm Research 26 (4):
35565.
Ye, R., D. G. Hayes, and R. Burton. 2014. Effects of particle size of
sucrose suspensions and pre-incubation of enzymes on lipase-cata-
lyzed synthesis of sucrose oleic acid esters. Journal of the American
Oil ChemistsSociety 91 (11):1891901. doi: 10.1007/s11746-014-
2537-8.
Ye, R., S.-H. Pyo, and D. G. Hayes. 2010. Lipase-catalyzed synthesis of
saccharide-fatty acid esters using suspensions of saccharide crystals
in solvent-free media. Journal of the American Oil ChemistsSociety
87 (3):28193. doi: 10.1007/s11746-009-1504-2.
Youngs, C. G. 1958. Preparation of esters and anhydrides from long
chain fatty acid chlorides. Journal of the American Oil Chemists
Society 35 (8):4167. doi: 10.1007/BF02632559.
Zhang, C.-Y., J.-G. Luo, R.-H. Liu, R. Lin, M.-H. Yang, and L.-Y.
Kong. 2017. Physakengoses K-Q, seven new sucrose esters from
Physalis alkekengi var. franchetii.Carbohydrate Research 449:1204.
doi: 10.1016/j.carres.2017.07.010.
Zhang, X., W. Wei, X. Cao, and F. Feng. 2015. Characterization of
enzymatically prepared sugar medium-chain fatty acid monoesters.
Journal of the Science of Food and Agriculture 95 (8):16317. doi: 10.
1002/jsfa.6863.
Zhang, X., F. Song, M. Taxipalati, W. Wei, and F. Feng. 2014.
Comparative study of surface-active properties and antimicrobial
activities of disaccharide monoesters. PLOS One 9 (12):e114845doi:
10.1371/journal.pone.0114845.
Zhao, L., H. Zhang, T. Hao, and S. Li. 2015. In vitro antibacterial activ-
ities and mechanism of sugar fatty acid esters against five food-
related bacteria. Food Chemistry 187:3707. doi: 10.1016/j.foodchem.
2015.04.108.
Zhao, R., Z. Chang, A. Jin, W. Li, B. Dong, and X. Miao. 2014.
Heterogeneous base catalytic transesterification synthesis of sucrose
ester and parallel reaction control. International Journal of Food
Science & Technology 49 (3):85460. doi: 10.1111/ijfs.12376.
Zhao, Y., Z. Chen, and T. Wu. 2018. Cryogelation of alginate improved
the freeze-thaw stability of oil-in-water emulsions. Carbohydrate
Polymers 198:2633. doi: 10.1016/j.carbpol.2018.06.013.
Zhong, X., J. Qian, M. Liu, and L. Ma. 2013. Candida rugosa lipase-cat-
alyzed synthesis of sucrose-6-acetate in a 2-butanol/buffer two-phase
system. Engineering in Life Sciences 13 (6):56371. doi: 10.1002/elsc.
201200170.
Zia-Ud-Din, H. Xiong, Z. Wang, L. Chen. and I. Ullah. 2019. Effects of
sucrose fatty acid ester addition on the structural, rheological and
retrogradation behavior of high amylose starch-based wood adhe-
sive. International Journal of Adhesion and Adhesives 89:518. doi:
10.1016/j.ijadhadh.2018.11.012.
Zia-Ud-Din, H. Xiong, Z. Wang, P. Fei, I. Ullah, A. B. Javaid, Y.
Wang, W. Jin, and L. Chen. 2017. Effects of sucrose fatty acid esters
on the stability and bonding performance of high amylose starch-
based wood adhesive. International Journal of Adhesion and
Adhesives 104:84653.
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 21
... Several investigations have studied the antibacterial action of carbohydrate ester derivatives, a phenomenon that can be utilized to maintain nutritional substances 90 . The mechanism of the antibacterial action was generally www.nature.com/scientificreports/ ...
... Figure 12 illustrates the relationship between the tittle esters and mean relative viability of (HePG-2) cell line (%) and Fig. 13 represent the effect of these esters on the in vitro IC 50 cancer cell lines (HePG-2). The mode of action of the possible cytotoxic route was believed to be disrupting the cancer cell membranes, which occurs in a detergency mechanism 90 . The antitumor effect has been suggested to be due to esters themselves, rather than their metabolites 99 . ...
Article
Full-text available
Sugar esters display surface-active properties, wetting, emulsifying, and other physicochemical phenomena following their amphipathic nature and recognize distinct biological activity. The development of nutritional pharmaceuticals and other applications remains of great interest. Herein, three novel homologous series of several N-mono-fatty acyl amino acid glucosyl esters were synthesized, and their physicochemical properties and biological activities were evaluated. The design and preparation of these esters were chemically performed via the reaction of glucose with different fatty acyl amino acids as renewable starting materials, with the suggestion that they would acquire functional characteristics superior and competitive to certain conventional surfactants. The synthesized products are characterized using FTIR, ¹H-NMR, and ¹³C-NMR spectroscopy. Further, their physicochemical properties, such as HLB, CMC, Γmax, γCMC, and Amin, were determined. Additionally, their antimicrobial and anticancer efficiency were assessed. The results indicate that the esters' molecular structure, including the acyl chain length and the type of amino acid, significantly influences their properties. The measured HLB ranged from 8.84 to 12.27, suggesting their use as oil/water emulsifiers, wetting, and cleansing agents. All esters demonstrate promising surface-active characteristics, with moderate to high foam production with good stability. Notably, compounds 6-O-(N-dodecanoyl, tetradecanoyl cysteine)-glucopyranose (34, 35), respectively and 6-O-(N-12-hydroxy-9-octadecenoyl cysteine)-glucopyranose (38) display superior foamability. Wetting efficiency increased with decreasing the chain length of the acyl group. The storage results reveal that increasing the fatty acyl hydrophobe length enhances the derived emulsion's stability for up to 63 days. Particularly, including cysteine in these glucosyl esters improves wetting, foaming, and emulsifying potentialities. Furthermore, the esters exhibit antibacterial activity against several tested Gram-positive and Gram-negative bacteria and fungi. On the other hand, they show significant antiproliferative effects on some liver tumor cell lines. For instance, compounds 6-O-(N-12-hydroxy-9-octadecenoylglycine)-glucopyranose (28), 6-O-(N-dodecanoyl, hexadecanoyl, 9-octadecenoyl and 12-hydroxy-9-octadecenoylvaline)- glucopyranose (29, 31, 32 and 33), respectively in addition to the dodecanoyl, hexadecanoyl, 9-octadecenoyl and 12-hydroxy-9-octadecenoyl cysteine glucopyranose (34, 36, 37 and 38), respectively significantly inhibit the examined cancer cells.
... Since sucrose esters can affect interfacial properties and fat crystallization, sucrose esters of different compositions have been widely used in food systems and industry (Nelen et al., 2014). Sucrose esters have attracted attention in the study of emulsions or whipping aerated systems due to their environmental compatibility and extensive hydrophilic-hydrophobic balance, and will also become the focus of future attention (HLB) values (Jiang et al., 2018;Teng et al., 2021). Sucrose esters affect shear-induced partial coalescence and surface-mediated partial coalescence (Fig. 1a) during whipping by synergistic and competitive adsorption at the oil/water interface, and the different mechanisms are shown in Fig. 3e. ...
... The competitive adsorption mechanism of sucrose esters and proteins at different concentrations is shown in Fig. 3b; when the sugar ester content exceeds 0.1225 wt%, an excessive amount of free surfactant will compete with the existing surfactant-protein complex, resulting in a decrease in the protein adsorbed around the droplet. However, in the presence of more sucrose esters (0.35 wt%), it will compete with the surfactant-protein complex and reduce the protein on the droplet surface, which will lead to complete a coalesce of the droplets during foaming, although bubbles will also form at this time, but with less stability and hardness (Teng et al., 2021). Zeng et al. (2021) found that low concentrations of sucrose ester S370 (< 0.05 wt%) resulted in reduced whipped cream overrun, and high concentrations (> 0.05 wt%) were the opposite. ...
Article
Full-text available
Whipped cream is popular among consumers due to its smooth texture and delectable taste. In recent years, it has also attracted increasing attention from producers and researchers and has been widely developed in the food industry. However, whipped cream systems are complex and unstable, and additives are applied to the system to obtain a more stable and excellent whipped cream product. Additionally, the high-fat content in whipped cream has a potential health risk to consumers, so the concept of low-fat whipped cream has been gradually developed as a strategy. However, the sensory and whipping properties of low-fat whipped cream need to be improved. This paper systematically discusses the effects of five types of additives (hydrocolloids, proteins, emulsifiers, carbohydrates, and lipid compounds) applied to improve the whipping, rheological, electrostatic, sensory, and nutritional properties of whipped cream. There are limited studies about nutritional properties, which should be the focus of future work, and the strategy of functional whipped cream is also a new direction. Further, this paper provides information on standards and regulations for additives, aiming to avoid the risks associated with the inappropriate application of additives by researchers and traders. The ultimate objective is to provide a healthy whipped cream with exceptional sensory properties and heightened nutritional value. Graphical Abstract
... Also, the moisture-retaining activity of Bangia fusco-purpurea extracts was better than that of the polysaccharides extracts of Saccharina japonica, Porphyra haitanensis, Bryopsis plumose, Codium fragile and Enteromorpha linza [40]. In fact, sugar esters are food additives recommended by the Food and Agriculture Organization of the United Nations and are widely used in food [41]. In Japan, sugar esters are widely used in the preservative application of canned foods [36]. ...
... Although silica gel column chromatography can process a large number of samples, it has some disadvantages of serious irreversible adsorption, low isolation efficiency and large solvent consumption, so it is necessary to find more advantageous resins to isolate glycolipids. After the initial isolation of the glycolipids from macroalgae, these glycolipids were further purified or prepared by Sephadex LH-20 gel column chromatography [5], or silica gel column chromatography [37,44], and/or preparative thin layer chromatography [11,41,45]. Finally, two glycolipids were isolated and purified from Bangia fuscopurpurea (Figure 9). ...
Article
Full-text available
In order to explore the extraction and activity of macroalge glycolipids, six macroalgae (Bangia fusco-purpurea, Gelidium amansii, Gloiopeltis furcata, Gracilariopsis lemaneiformis, Gracilaria sp. and Pyropia yezoensis) glycolipids were extracted with five different solvents firstly. Considering the yield and glycolipids concentration of extracts, Bangia fusco-purpurea, Gracilaria sp. and Pyropia yezoensis were selected from six species of marine macroalgae as the raw materials for the extraction of glycolipids. The effects of the volume score of methanol, solid–liquid ratio, extraction temperature, extraction time and ultrasonic power on the yield and glycolipids concentration of extracts of the above three macroalgae were analyzed through a series of single-factor experiments. By analyzing the antioxidant activity in vitro, moisture absorption and moisturizing activity, the extraction process of Bangia fusco-purpurea glycolipids was further optimized by response surface method to obtain suitable conditions for glycolipid extraction (solid-liquid ratio of 1:27 g/mL, extraction temperature of 48 °C, extraction time of 98 min and ultrasonic power of 450 W). Bangia fusco-purpurea extracts exhibited a certain scavenging effect on DPPH free radicals, as well as good moisture-absorption and moisture retaining activities. Two glycolipids were isolated from Bangia fusco-purpurea by liquid–liquid extraction, silica gel column chromatography and thin-layer chromatography, and they showed good scavenging activities against DPPH free radicals and total antioxidant capacity. Their scavenging activities against DPPH free radicals were about 60% at 1600 µg/mL, and total antioxidant capacity was better than that of Trolox. Among them, the moisturizing activity of a glycolipid was close to that of sorbierite and sodium alginate. These two glycolipids exhibited big application potential as food humectants and antioxidants.
... Saponification value of the tested oil samples was determined 16 . Briefly, 5 g of the tested sample was weighted into 250-300 ml conical flask, and 50 ml alcoholic KOH solution (35)(36)(37)(38)(39)(40) g KOH were dissolved in 20 ml water and diluted to one liter with alcohol, 95%) was added. Flask was connected with air condenser, boiled until fat was completely saponified (~ 30 min), cooled and titrated with 0.5 M HCl using phenolphthalein. ...
... Sucrose stearate showed antimicrobial activities against some gram positive and some gram-negative bacteria 36,37 . This activity was related to the sugar moiety which rupture the bacterial cell membrane 38 . ...
Article
Full-text available
Food emulsifier are mostly prepared from a lipophilic lipid tail with a hydrophilic sugar head. In this study, the lipophilic tail was obtained from apricot kernels, which are food waste, and the hydrophilic head was gluconic acid instead of sugar, in order to draw attention to the non-cyclic poly hydroxyl compounds. Thus, oleic acid of apricot kernel was used as the lipophilic moiety of the prepared surfactant. So, apricot kernel was grinned and dried, oil was extracted using soxhlet apparatus, Physical and chemical parameters and fatty acids composition of the extracted oil had been determined. The extracted oil was then hydrolyzed into glycerol and a mixture of free fatty acids. The fatty acids mixture was separated. Then, oleic acid was extracted individually in pure form using supercritical CO2 extractor, it was then confirmed according to its melting point, Gas chromatography–mass spectrometry (GC–MS) after esterification, elemental analysis, Proton nuclear magnetic resonance (H¹NMR), and mass spectrometry (MS) to detect the corresponding molecular ion peak. The pure individual oleic acid was converted to hydroxy stearic acid, which was then converted to an amphiphilic compound (surfactant) via esterification reaction with the hydrophilic gluconic acid, and afforded a new surfactant known as 2,3,4,5-tetrahydroxy-6-((9-((-2,3,4,5,6-pentahydroxyhexanoyl) oxy)octadecanoyl) oxy)hexanoic acid or stearyl gluconate for simplification. The structures elucidation of all synthesized compound was established according to elemental analysis and spectral data (Fourier transform infrared IR, ¹H NMR, ¹³C NMR and MS). Moreover, the prepared compound was tasted for its antibacterial activity, and showed good activities against some types of bacteria. The surface-active properties, foamability, foaming stability and emulsion stability of stearyl gluconate were studied and compared with the properties of the well-known surfactant sucrose stearate, and it was clear that, the activity of stearyl gluconate as a surfactant was higher than that of sucrose stearate. Moreover, establishment of safety of this compound was performed using albino rats by acute oral toxicity and kidney and liver functions of these mice. On the other hand, the prepared surfactant was used in the production of low fat—free cholesterol mayonnaise as egg replacer. Texture properties and the sensory evaluation of the prepared mayonnaise showed that the properties were improved by using the new prepared surfactant. Thus, the prepared gluconyl stearate can be used as a safe food additive.
... SFAE has triggered enormous interest in discovering new applications for these materials, such as antimicrobial [4][5][6][7], anticancer [8], insecticidal activities [9], and drug delivery [10][11][12][13][14]. There is no doubt that commercial sources of SFAE are available for various applications [15]. ...
... This clearly indicated the more compactness of the lauroyl (C12) chain than the decanoyl (C10) and stearoyl (C18) chains. 15 ...
Preprint
Full-text available
The approval of sucrose fatty acid esters (SFAEs) as food additives/preservatives has triggered enormous interest in discovering new applications for these materials. Accordingly, many researchers reported that SFAEs consist of various sugar moieties, and hydrophobic side chains are highly active against certain fungal species. The combination of chain length and site of acylation is crucial in endowing the SFAE with high antimicrobial potency against Aspergillus species. Following several important studies, we herein present the synthesis and an assessment of the effects of acylation site and chain length (i.e., C-6 vs. C-2, C-3, C-4, and long-chain vs. short-chain) on the antimicrobial activity of mannopyranoside fatty acid esters. In vitro tests revealed that the fatty acid chain length in mannopyranoside esters significantly affects the antifungal activity where C12 chains are more potent against Aspergillus species. In terms of acylation site, mannopyranoside esters with a C8 chain substituted at the C-6 position are more active in antifungal inhibition. Molecular docking also revealed that these mannopyranoside esters had comparatively better stable binding energy, and hence better inhibition, with the fungal enzymes lanosterol 14-alpha-demethylase (3LD6), urate oxidase (1R51), and glucoamylase (1KUL) than the standard antifungal drug fluconazole. Additionally, the thermodynamic, orbital, drug-likeness, and safety profiles of these mannopyranoside esters were calculated and discussed, along with the structure-activity relationships (SAR). This study thus highlights the importance of the acylation site and lipid-like fatty acid chain length that govern the antimicrobial activity of mannopyranoside-based SFAE.
Article
Full-text available
Palm oil (PO), a semi-solid fat at room temperature, is a popular food ingredient. To steer the fat functionality, sucrose esters (SEs) are often used as food additives. Many SEs exist, varying in their hydrophilic-to-lipophilic balance (HLB), making them suitable for various food and non-food applications. In this study, a stearic–palmitic sucrose ester with a moderate HLB (6) was studied. It was found that the SE exhibited a complex thermal behavior consistent with smectic liquid crystals (type A). Small-angle X-ray scattering revealed that the mono- and poly-esters of the SE have different packings, more specifically, double and single chain-length packing. The polymorphism encountered upon crystallization was repeatable during successive heating and cooling cycles. After studying the pure SE, it was added to palm oil, and the crystallization behavior of the mixture was compared to that of pure palm oil. The crystallization conditions were varied by applying cooling at 20 °C/min (fast) and 1 °C/min (slow) to 0 °C, 20 °C or 25 °C. The samples were followed for one hour of isothermal time. Differential scanning calorimetry (DSC) showed that nucleation and polymorphic transitions were accelerated. Wide-angle X-ray scattering (WAXS) unraveled that the α-to-β′ polymorphic transition remained present upon the addition of the SE. SAXS showed that the addition of the SE at 0.5 wt% did not significantly change the double chain-length packing of palm oil, but it decreased the domain size when cooling in a fast manner. Ultra-small-angle X-ray scattering (USAXS) revealed that the addition of the SE created smaller crystal nanoplatelets (CNPs). The microstructure of the fat crystal network was visualized by means of polarized light microscopy (PLM) and cryo-scanning electron microscopy (cryo-SEM). The addition of the SE created a finer and space-filling network without the visibility of separate floc structures.
Article
Oleogels have been explored as fat substitutes due to their healthier composition compared to trans and saturated fats, also presenting interesting technological perspectives. The aim of this study was to investigate the compositional perspective of multicomponent oleogels. Structuring ability of lecithin (LEC) (20 or 90 wt% of phosphatidylcholine – PC) combined with glycerol monostearate (GMS), sorbitan monostearate (SMS) or sucrose monostearate (SAC) in sunflower oil was evaluated from oleogels properties. The thermal and rheological properties, microstructure and stability of the oleogels were affected by the difference in the chemical composition of LEC and the ratio between LEC and different surfactants. Interestingly, low-phosphatidylcholine LEC (L20) performed better, although systems formed with reduced amounts of LEC tended to be softer (LEC-GMS) and present high oil holding capacity (LEC-SMS). The mixtures of LEC and monostearate-based surfactants showed different behaviors, depending on the surfactant polar head. In LEC-GMS systems, LEC hindered the self-assembly of GMS in sunflower oil, compromising mechanical properties and increasing oil release. When combined with SMS, LEC acted as a crystal habit modifier of SMS, forming a more homogeneous microstructure and producing stronger oleogels with greater oil binding capacity. However, above the threshold concentration, LEC prevented SMS self-assembly, resulting in a weaker gel. A positive interaction was found in LEC-SAC formulations in specific ratios, since SAC cannot act as a single oleogelator. Results show the impact of solubility balance played by LEC and fatty-acid derivatives surfactant when combined and used as oleogelators. This knowledge can contribute to a rational perspective in the preparation and modulation of the properties of edible oleogels.
Article
Full-text available
The amino acid composition of subtilisin Carlsberg has been determined and is in good agreement with the 274 residues found by sequence studies in the protein. From a tryptic hydrolysate, it was possible to resolve the peptides into a series of fractions by gradient elution from Dowex 50. After further purification, there were obtained the 14 tryptic peptides, accounting for the complete composition of the protein. In addition, some peptides resulting from secondary cleavage were isolated. A peptide containing a bond susceptible to trypsin was obtained in pure form. The compositions and properties of these peptides are reported.
Article
Full-text available
Enzymes are natural catalysts highly specific to the substrate type and operate under mild conditions of temperature, pressure, and pH with high conversion rates, which makes them more efficient than conventional chemical catalysts. The enzymes can be obtained from various sources, animal, vegetable, and microbiological. Lipases are very versatile enzymes, and this has aroused the interest of the industries. However, the great problem of the use of soluble lipases is the high cost of acquisition, low operational stability, and difficulties of recovery, and reuse. Enzymatic immobilization has been suggested as an alternative to reduce the limitations of soluble enzymes, increasing their stability and facilitating recovery, and reuse, significantly reducing the cost of processes involving the use of enzymes. This review presents a discussion on the different immobilization methods for lipase, as well as the challenges of use lipases immobilized on the industrial scale.
Article
“The Structure and Properties of Ionic Melts” was the title of a Faraday Society Discussion held in Liverpool in 1961; it dealt exclusively with molten inorganic salts. “Ionic Liquids” was the title of Chapter 6 of the textbook Modern Electrochemistry by Bockris and Reddy, published in 1970: it discussed liquids ranging from alkali silicates and halides to tetraalkylammonium salts.
Article
The moisture stability of tobacco shred, a typical porous carbohydrate material, is very important during its processing, storage and smoking, moreover, it is sensitive to environmental conditions. Therefore, effect of sucrose esters (SEs) and sucrose ester/Polygonatum sibiricum polysaccharide mixture (SPMs) on the moisture retention and moisture resistance of tobacco shred was assessed. When SEs were added to tobacco shred, moisture resistance was significantly enhanced, whereas moisture holding capacity was attenuated. Contrarily, the addition of SPMs made moisture retention index (MRI) and moisture proof index (MPI) increase from 1.8910 to 2.1612 and from 1.9489 to 2.0665, respectively, revealing that SPMs improved the moisture retention and moisture proof ability of tobacco shred simultaneously. The monolayer moisture content (M0) was decreased by SEs and increased by SPMs. Low-field nuclear magnetic resonance (LF-NMR) analysis showed that during adsorption, SPMs reduced the interaction between tobacco shred and water via hydrophobic property of SEs; during desorption, SPMs promoted the interaction between tobacco shred and water through hydrophilic binding of polysaccharide, leading to the migration of immobilized water to bound state. The modeling of the isotherms and LF-NMR analysis clarified the mechanism why SPMs could improve moisture stability of tobacco.
Article
Pancakes served as a dessert with creams and fruits toppings nowadays stored under refrigerated conditions are often subjected to significant changes in physical properties recognized as quality deterioration, which is efficiently prevented by including small molecule emulsifiers into commercial pancakes formulations. In order to reveal these molecular prevention mechanisms in the real multicomponent wheat-based products, small molecule emulsifiers with various HLB values and head group size were formulated into pancake samples and their physical properties evaluated. Starch retrogradation, emulsifier-starch inclusion complex formation and gelatinization behaviors were followed by the texture profile analysis, differential scanning calorimetry, iodine colorimetry and confocal laser scanning microscopy. Hardening of the pancake samples varied with the type of emulsifiers, which in general reduced the retrogradation of starch. This reduction was dominantly attributed to the emulsifier-starch complex formation, particularly with amylose molecules, probably leading to less leaching of the starch molecules from the starch granules. These findings are of practical use in selecting effective emulsifiers that prevent hardening of pancakes. The degree of complexation was partially explained by the HLB and head group size of the emulsifiers, while it can be also significantly affected by interactions between the emulsifier and other components in the pancake mixture, like milk and eggs.
Article
Hypothesis: The development of functional and nutritional surfactants for the food industry remains a subject of great interest. Herein, therefore, we report on the design and synthesis of novel trisaccharide (raffinose) monoester-based surfactants in the expectation that they would display functional properties superior to certain disaccharide-based, commercially-deployed emulsifiers and thus have potential for industrial applications. Experiments: The title esters were prepared by enzymatic methods and their properties as surfactants evaluated through determination of their HLB values, water solubilities, CMCs, foamabilities and foaming stabilities as well as through investigation of their impacts on the stability of oil-in-water emulsions over a range of storage times and under certain other conditions. Findings: The emulsifying properties of 6-O-acylraffinose esters are dictated, in large part, by the length of the associated alkyl chains. The results of storage and environmental stress experiments revealed that the increasing length of alkyl chains enhances the stability of the derived emulsions. All the raffinose ester-stabilized oil-in-water emulsions displayed stratification effects under strongly acidic conditions (pH ≤ 4) or at high ionic strength (≥300 mM) while possessing reasonable resistance to variations in temperature. As such, a number of the raffinose monoesters showed greater stability to environmental stress than their commercially-deployed and sucrose-based counterparts. The structure-property profiles established through the present study provide a definitive guide for the development of raffinose esters as novel emulsifiers, particularly in the food industry.
Article
Ionic liquids (ILs), are considered as eco-friendly solvents due to their low vapor pressure and non-volatile properties compared to organic solvents and can be employed during processes for the extraction and/or conversion of lignin and chitin to value-added products which can improve the overall economics of the integrated biorefinery concept. However, relatively few studies have investigated their ability to solubilize and modify during solvation of lignin and chitin in terms of crystallinity, molecular weight, etc., with a significant focus on polysaccharides like cellulose and hemicellulose. Similarly, the biomedical applications of lignin and chitin are not recognised so much as compared to polysaccharides. The aim of this review is survey the current literature based on the cutting-edge-knowledge regarding the solubilization of chitin and lignin in various classes of ILs and furthermore, the upgradation of these biopolymers and/or their derivatives for the biomedical applications. Moreover, this review provides a comprehensive summary of the latest progress on the solubilization of integrated biorefinery lignin and chitin with a wide range of ILs using experimental and theoretical studies.
Article
This work studied the ultrasound technology as an alternative to enhance the reaction of sucrose inversion. To achieve it, the effect of ultrasound processing was evaluated in three situations: the reaction (invertase acting on substrate under ultrasound processing), the isolated substrate (sucrose) and the isolated enzyme (invertase). Ultrasound (25 kHz, 22 W/L) processing was conducted under different conditions of pH and temperature. The ultrasound assisted reaction was accelerated in relation to the conventional processing. This process increased the sucrose hydrolysis rate up to 33% at 40 °C, and 30% at 30 °C. The ultrasound as a pre-treatment method did not promote sucrose hydrolysis, but slightly reduced the invertase activity, independently of the pH and temperature of evaluation. Sonication also increased Vmax (increased of 23%) and maintained constant Km, indicating that the ultrasound accelerated mass transfer during the reaction, but did not directly affected the enzyme. In addition, the invertase catalytic efficiency enhanced 27% under sonication. Therefore, ultrasound technology emerges as an interesting alternative to improve the invertase performance, accelerating enzymatic reaction.
Article
In this study, sucrose fatty acid ester (SE) was used to evaluate the characteristics of a high amylose starch-based wood adhesive. Fourier transform infrared (FT-IR) spectroscopy confirmed the occurrence of graft copolymerization reactions whereas X-rays diffraction (XRD) analysis verified the presence of amylose-SE complexes. Scanning electron microscopy (SEM) a revealed uniform distribution of spaces whereas transmission electron microscopy (TEM) showed good dispersion of latex particles with SE addition as evident by small and poly-dispersed particles in the wood adhesive. The water dynamics in the starch adhesive system was studied using low-field nuclear magnetic resonance (LF-NMR). Furthermore, the addition of SEs resulted in enhanced shear-thinning and solid like-behaviors and anti-retrogradation properties of the adhesive with incorporation of SE. These results showed that SEs could improve the rheological and anti-retrogradation properties of the wood adhesive and offers a major step forward to prepare bio-based adhesives as an alternative for petroleum-based wood adhesives.