Você está na página 1de 43

Accepted Manuscript

Kinetics of the transesterification of methyl palmitate and sucrose using surfac-


tants

María F. Gutiérrez, Jose L. Rivera, Andrea Suaza, Alvaro Orjuela

PII: S1385-8947(18)30656-9
DOI: https://doi.org/10.1016/j.cej.2018.04.085
Reference: CEJ 18885

To appear in: Chemical Engineering Journal

Received Date: 21 September 2017


Revised Date: 5 April 2018
Accepted Date: 13 April 2018

Please cite this article as: M.F. Gutiérrez, J.L. Rivera, A. Suaza, A. Orjuela, Kinetics of the transesterification of
methyl palmitate and sucrose using surfactants, Chemical Engineering Journal (2018), doi: https://doi.org/10.1016/
j.cej.2018.04.085

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Kinetics of the transesterification of methyl palmitate and sucrose using surfactants

María F. Gutiérrez, Jose L. Rivera, Andrea Suaza, Alvaro Orjuela 1

Department of Chemical and Environmental Engineering, Universidad Nacional de

Colombia, 111321 Bogotá D.C., Colombia.

Abstract

This work studied the synthesis of sucroesters via transesterification of sucrose and

methyl palmitate in a solvent-free process. The kinetic experiments were carried out under

slight vacuum (80 kPa) to ensure an irreversible reaction, with K2CO3 as catalyst, and

using a variety of surfactants (potassium palmitate, sucrose palmitate and glyceryl

monostearate) with different Hydrophilic-Lipophilic Balance (HLB) as contacting agents.

Reaction samples were analyzed with a novel isocratic HPLC method using a refraction

index detector. The kinetic effects of temperature (369-409 K) and surfactant type and

concentration (5 to 15%wt) were assessed. Results revealed that the reaction exhibited a

lag period, consistent with a limiting sucrose activation step, followed by faster

transesterification to monoester. Also, the different substitution steps on the sucrose

backbone showed strong temperature dependence, and the corresponding large energies

of activation. According to the product distribution, the reaction proceeded mostly to mono,

di and tri substituted sucroesters. Particularly, potassium palmitate exhibited synergic

contact-enhancing and catalytic effects, delivering a sucrose ester productivity of 55 g

SE/h kg and a selectivity to monoesters in the range of 52 to 67%. It was observed that

1
*Corresponding author. Tel (+571) 3165000 x 14303. E-mail address: aorjuelal@unal.edu.co
high HLB surfactants promoted a higher selectivity for monosubstituted sucroesters. The

reaction kinetics was modeled as a set of steps involving sucrose activation, monoester

formation, and lumped oligoester synthesis. The kinetic parameters were regressed for

each surfactant, and the model showed good agreement with experiments. Obtained

models can be used for further process analysis and design.

Keywords

Sucrose ester, Kinetics, Transesterification, Surfactants, Methyl palmitate


1. Introduction

Sucrose esters are entirely-biobased nonionic molecules used as surfactants in niche

applications such as cosmetics, food and feed (ingredient number E473), clean and care

products, and pharmaceuticals. Nowadays these esters represent a small portion of the

nonionic surfactants market, that is mostly dominated by poly-hydroxylated molecules

(alcohols, alkyl phenols, amines, thiols, etc.) obtained by poly-addition of ethylene or

propylene oxides [1]. Unlike most polyoxyethylene structures, sucroesters show interesting

properties as green chemical such as rapid biodegradability, biocompatibility, and biocide

potential for certain microorganisms [2–5]. These features have revitalized sucrose esters

demand, as they are considered ecofriendly and more sustainable than the traditional

nonionic surfactants.

Biobased sucrose esters are usually produced via base-catalyzed or enzymatic

transesterification of sucrose with fatty acids methyl esters (FAMEs) or vinyl esters

(FAVEs). Both are equilibrium reactions, thus the removal of the produced alcohol

(methanol or vinyl alcohol) helps shifting the reaction toward the desired products (Figure

1). Despite transesterification reactions are usually straightforward, one of the main

challenges in sucroesters production is to overcome the incompatibility between reactants.

Sucrose is a highly polar solid and FAMEs are highly non-polar viscous liquids, which

results in poor contact between reactants, low reaction rates and limited conversions.

Although a temperature increase would be an enhancing factor, the thermal instability of

reactants and products, limits the maximum allowed temperature in the process.

As an alternative to improve compatibility, the reaction has been carried out within solvents

of medium polarity such as dimethylsulfoxide (DMSO) or dimethylformamide (DMF), in


which both reactants are soluble [6,7]. However, the required solvent loading to achieve

complete solubility of reactants is around 60%wt of the reaction mixture; therefore, the

reaction productivity is very low. In addition, an exhaustive solvent removal in the final

product is necessary owing to the final use of sucrose esters and the health concerns of

DMF and DMSO [8–10]. This represents a major separation challenge because of the high

boiling temperature of the solvents (i.e. 426 K for DMF, and 462 K for DMSO) and the

thermal instability of the biobased molecules.

In this direction, solvent-free base-catalyzed transesterification has shown to be a valuable

alternative during sucroesters production. In this process, a surfactant is used as

contacting agent that helps dispersing/emulsifying sucrose in FAMEs. Alkaline and

multivalent fatty acid soaps have demonstrated a good emulsifying action for the process

[11–15]. Besides, sucroesters have been also used to improve reactants compatibility [16],

indicating that the reaction rate is improved as the reaction proceeds [14]. Mono and diacyl

glycerol esters have been also evaluated as surfactants, producing a final mixture of

sucroesters and sucrose glycerides (food ingredient E474) [17]. Surfactants loading has

been evaluated in the range of 1 to 30%wt, at temperatures from 403 to 453K [15,18].

Regarding catalysis, potassium carbonate (K2CO3) is the most widely used catalyst for this

reaction, at concentrations between 5% and 12%wt [19]. The molar feed ratio of sucrose

to FAME has been evaluated in the range of 0.1 to 2, in order to tune the degree of

substitution in the sucrose backbone [13].

The reaction product can be a complex mixture of esters with a different degree of

substitution because there are multiple reactive hydroxyl groups in sucrose (Figure 1).

Commercial sucroesters mostly contain mono and diesters (no less than 80 %wt), while

the mixtures with a higher degree of substitution (up to octa-esters) are commercialized as
type I or type II sucrose oligoesters. The commercial differentiation of these products lies

in the fact that the substitution degree determines the properties as surfactants and their

final applications [4]. The Hidrofilic-Lipofilic Balance (HLB) is an index ranging from 0 to 20

that indicates which kind of emulsions tends to form a specific surfactant. While sucrose

monoesters are water soluble and tend to form oil-in water emulsions (high HLB), highly

substituted sucrose esters are lipophylic molecules used for water-in-oil emulsions (low

HLB) or as fat substitutes [20, 21]. Generally, the substitution degree depends on the

operating conditions during synthesis such as temperature, catalyst nature and

concentration, reactants molar ratio, the use of additives (e. g. solvents, surfactants), etc.

[20,22]. Therefore, the capability of tuning the process conditions to obtain a specific

product is of major importance.

Figure 1. Sucrose fatty acid esters production by transesterification of fatty acid methyl

ester (FAME) and sucrose


In addition to the challenge of tuning the substitution degree, transesterification can occur

in any of hydroxyl groups, because all are at some extent reactive. Despite the most

favorable positions are the 6, 6’ and 1’ carbons in the sucrose backbone (Figure 1),

substitution can occur in any hydroxyl group [23]. This means that a sucrose monoester is

in fact a mixture of structural isomers, in which the fatty acid chain is bonded to different

hydroxyl groups. The structural isomers are also formed in products with higher degree of

substitution. An additional complication appears if a mixture of FAMEs (e.g biodiesel) is

used as reactant instead of a single methyl ester. Here, the length of the fatty acid and its

degree of unsaturation will play an important role in the properties of the final sucroesters

mixture.

The reaction kinetics of the solvent process has been fairly studied, using DMF and

DMSO, under different conditions [24]. In the other hand, and despite the solvent-less

process was first reported in 1977 [19], it was just recently explored using mono- and

divalent soaps as surfactants [11]. Among the different surfactant reported as contacting

agents, those used in this work were selected taking into account green chemistry and

process integration principles. When using sucroesters as surfactant, further separation

processes of the contacting agent are not needed, reducing purification steps. On the

other hand, both, potassium palmitate (i.e. potassium soap) and glycerol monostearate are

byproducts during FAMEs production [25], and they could be present in a crude feedstock

(less expensive) [26]. Moreover, these biobased surfactants could also be present in the

product depending on the final application. A mixture of potassium palmitate and sucrose

esters can be used in cleaning applications. In a similar way, a mixture of monoglycerides

and sucrose esters can be used in food and personal care products as surfactants, as

both are edible and mild surfactants (e.g. SUCRAMULSE®).


Considering all the above, this work developed a kinetic study of the transesterification of

sucrose and methyl palmitate using potassium palmitate (KP), glycerol monostearate

(GMS) and sucrose palmitate (SE) as contacting agents. These surfactants have different

affinities: potassium soap is a water soluble ionic surfactant (HLB~20), glycerol

monostearate is an oil soluble nonionic surfactant (HLB~4), and commercial sucroester is

a water soluble nonionic surfactant (HLB~15). The effect of temperature and surfactant’s

concentration on the reaction rate and selectivity was evaluated. Kinetic models for each

surfactant were obtained, and they can be used for further process up-scaling, tuning the

selectivity to a desired product.

2. Materials and methods

2.1. Materials

The list of reactants, catalysts, surfactants and analytical standards during experiments is

presented in Table 1. These chemicals were used without further purification.

Table 1. Chemicals and reagents used in experiments

Chemical Grade Purity Company


Methyl palmitate Analysis 99.5% Sigma Aldrich Co (USA)
Analytical standard 99%
Sucrose Analysis 99.9% Panreac AppliChem GmbH (Germany)
Potassium Palmitate (KP) Analysis 99.2% Wako Pure Chemical Ind. Ltd. (Japan)
Sucrose Palmitate (SE) Analytical standard 90% Alfa Aesar (USA)
FG - Modernist Pantry LLC (USA)
Glycerol monostearate (GMS) Analysis >95.0% Alfa Aesar (USA)
Potassium carbonate Analysis 99.74% Alfa Aesar (USA)
Acetronitrile (ACN) UHPLC 99.9% Panreac ApplyChem GmbH (Germany)
Dimethylsulfoxide (DMSO) Analysis 99.9% Panreac ApplyChem GmbH (Germany)
2.2. Methods

2.2.1. Kinetic experiments setup and procedures

In order to reduce difficulties during the chemical characterization of samples from

reactions, a single methyl ester was used as reactant for kinetic experiments. Methyl

palmitate was selected as a convenient model molecule for the kinetic experiments,

because this work is aimed to develop a process for palm-oil derived sucroesters. The

range of operating conditions were defined according to the aforementioned literature

reports. Molar ratio of methyl palmitate to sucrose was kept constant at 2.5. Potassium

carbonate (K2CO3, 5 %wt) was used as catalyst, and the surfactant concentrations ranged

from 5 to 15 %wt. The mixture of methyl palmitate and catalyst was homogenized using a

high-performance dispersing mixer (T-25 Ultraturrax®, IKA, Germany) at 16000 rpm and a

temperature of 320 K for 5 min before reaction. Then the contact agent was added and

homogenized for another 5 min. Once FAME catalyst and surfactant were mixed, a

defined amount of previously grinded sucrose (particle size ca. 125 µm) was added to the

mixture and homogenized for 10 min. This homogenization represented a further decrease

of sucrose particle size (~ 100 µm) without introducing water into the system. In addition,

during this process, the particle size of the catalyst was also reduced (ca. 11 ± 2 µm). The

characterization of the particle size of the solids was assessed by optical microscopy. The

homogenized mixture was placed in a 150mL glass-jacketed reactor, degasified at an

absolute pressure of 33.4 kPa, and kept under agitation during reaction at 700 rpm with a

cross shape magnetic stirrer. Reactions were run under isothermal conditions, and the

temperature was controlled with a thermal circulation bath (Polystat, Cole Parmer). The

produced methanol was recovered using a total condenser and a side volumetric collection

vessel. The temperature of the condenser was adjusted at 268 K using circulation bath

(Julabo MC, Julabo) operating with water-propylenglycol as refrigerant. The experimental


set up is showed in Figure 2. Samples of reaction (ca. 0.5 mL) were withdrawn at different

time intervals during the reaction period, using a pre-heated plastic 3mL pipette.

Previous kinetic reports for solvent-less sucroester synthesis missed to study the influence

of temperature, and the type and concentration of surfactant. In this work, a set of

experiments were performed to evaluate the effect of such variables in reaction. The

studied surfactants were potassium palmitate (KP), glycerol monostearate (GMS) and

sucrose ester (SE). For each surfactant, two factors at three levels were considered:

reaction temperature (369, 389 and 409 K) and surfactant’s concentration referred to the

total mass of sucrose and FAME (5, 10, and 15%wt). The set of experimental conditions

was stablished according to the factorial design presented in the Figure S-1 included in the

supplementary material. The complete set of conditions in the experimental design are

summarized in Table 2.

Methanol
condenser

Jacketed glass
reactor Methanol
volumetric
collector Refrigeration
bath

Heating
bath Magnetic
stirring plate

Figure 2. Experimental set-up for kinetic experiments

Preliminary tests were carried out to rule out mass transfer limitations. Sucrose was

grinded, sieved, and separated in fractions of different mesh sizes. The corresponding
mean particle sizes of the crystals were 302.5, 165, 127.5, 97.5, and 75 µm. In each

preliminary experiment, the reacting mixture was prepared with a 2.5 molar ratio of methyl

palmitate and sucrose with the specific mean particle size, potassium carbonate as

catalyst (5 %wt), and potassium palmitate as surfactant (5 %wt). In these exploratory

experiments, the pre-homogenization of the mixture with the Ultraturrax® was carried out

without sucrose, because the high shearing effect could change the crystals particle sizes.

Additional preliminary tests were performed to rule out external mass transfer limitations

during kinetic experiments by changing stirring rates. The range of operating conditions

were stablished taking into account the need for a proper suspension of the solids in the

liquid FAME. Reactions were performed as above described using stirring rates ranging

from 400 to 800 rpm. This range was limited by the solids loading in the mixture and

geometry of the reactor/stirrer. At lower stirring rates some it was not possible to suspend

the solids, and at higher stirring rates a strong vortex was observed. No differences in the

reaction rate were detected in this range of operating conditions.

2.2.2. Analysis

A novel quantification method was developed by using an isocratic HPLC method. Methyl

palmitate and sucrose esters with different degrees of esterification were quantified by

mean of a Dionex UltiMate 3000 (Thermo Scientific, USA) HPLC, equipped with a Shodex

RI-101 detector (Showa Denko K.K., Japan) and using a LiChrospher100 RP-18 5.0 µm

column (125 x 4.0 mm Merck, Germany). The system operated at 313.15K using an

isocratic elution with ACN/DMSO 80/20 v/v as mobile phase with a flow of 0.5mL/min. For

the analysis, weighted samples withdrawn from the reactor were dissolved (ca. 50mg/mL)

in a mixture of ACN/DMSO 60/40 v/v, and filtered in a 0.45µm nylon filter (Membrane

Solutions). The injection volume was 20µL. The Chromeleon 7.1 software was used for

data acquisition and processing. Calibration was performed by HPLC analysis of samples
of known compositions prepared by dilution of standards. Sucrose esters and methyl

palmitate quantification was possible above the limit of detection (4 ng). A linear response

was obtained in the whole concentration range, and a concentration uncertainty of ±

1.2%wt was observed in all samples. Quantification of sucrose with the developed

technique was not possible because of the limited solubility in the solvent, and the lack of

resolution in chromatograms. Standard components used in the sucrose esters

quantification were obtained by flash column chromatography by separating a commercial

sample of sucrose palmitate. The column was filled with bulk reversed phase C18 Flash

silica gel (40-60µm) and a gradient flow of MeOH/Water 80/20 to 100/100 in 45 min

followed by an isocratic flow of MeOH for 11 min were used to elute sucrose palmitate with

different esterification degrees. The chemical structures of the isolated components were

confirmed by FTIR and 13C-NMR.

3. Results and Discussion

A major challenge when studying sucrose ester synthesis is the quantitative analysis of

samples. First, as most molecules involved in the reaction are not chromophores, widely

available UV or diode array HPLC detectors are of limited use. Second, the difference in

polarities of reactants and products require suitable solvent mixtures to ensure complete

dissolution of samples. Additionally, previous studies on the sucrose ester synthesis

indicated that separation should be done by HPLC under gradient elution (non-isocratic

conditions) or by changing the mobile phase during separation [27]. These operating

conditions prevent from using a typical general-purpose refractive index detector because

the base line fluctuates with the change of nature and concentration of mobile phase.

Recent reports on sucroesters synthesis indicated a successful implementation of HPLC

separations using near-universal detectors such as the evaporative light scattering (ELSD)
or the Charged aerosol (CAD) detectors. However, they were not available in our facilities.

For this reason, we developed a successful isocratic HPLC separation technique using a

refractive index detector as described in section 2.3. A typical chromatogram of the

reaction products is presented in Figure S-2 in Supplementary Material. As described

above, all peaks were identified according to standards obtained by flash chromatography

of a commercial mixture of sucrose palmitates. Peak resolution was good and allowed a

reproducible quantification of the components. However, separation of the different

structural isomer esters was not possible, and the reported data correspond to the lumped

concentration of mono-esters, or di-esters, or tri-esters, and so on.

As mentioned, substitution can occur in different hydroxyl groups of the sucrose molecule.

Despite the initial molar excess of methyl palmitate with respect to sucrose used in

reactions (2.5 to 1), the relative ratio of functional groups of sucrose (equivalents of

hydroxyl groups) with respect to carboxyl groups of methyl palmitate is 8:1. This means

that the limiting reactant was methyl palmitate, and the conversion was evaluated with

respect to this component. As sucrose was not quantified using the reported HPLC

method, its concentration was calculated based upon the measured concentration of the

produced sucroesters. By neglecting oligomerization of sucrose, it can be assumed that

the saccharide moiety concentration is constant and equal to that fed at the beginning of

reaction (at  ). So, at any time () of reaction, the total concentration of saccharides is the

sum of sucrose ( ( plus monoesters ( ( and oligoesters ( (). Then, as

sucroesters were quantified by HPLC, sucrose concentration (mole fraction) was

computed by difference from the initial saccharides loading, using Equation 1.

 ( =  (  +  (  +  (  −  ( −  ( (1)


The total change in methyl palmitate molar concentration from the beginning ( ( ) to the

end of the reaction ( ( ) was used to calculate conversion of methyl palmitate ( ) as

shown in Equation 2. This balance was also used to verify the material balance according

with the distribution of sucroesters.

 (   


 = (2)
 ( 

The total productivity of reaction (, g/kg/h) was calculated as the mass of sucrose ester

per time per unit mass of reactive mixture. It was calculated from the initial and final mass

concentration of sucroesters, as shown in Equation 3.

     (  ( 


= ∗ 1000 (3)


3.1. Mass-Transfer considerations on the reaction

According to the reported reaction mechanism for the base catalyzed transesterification

[28,29], the interaction between the sucrose molecule and the alkaline catalyst results in

the activation of one of the eight hydroxyl groups in sucrose. Then, the formed alkoxide

attacks the carbonyl group of FAME and forms an intermediate molecule. This last

intermediate component generates the sucrose ester, releasing methanol. The solvent-

free production implies the occurrence of a solid-solid-liquid interaction among the catalyst,

sucrose and FAME, respectively. Because the phase splitting, two approaches have been

used to explain the interaction between sucrose and the catalyst that leads to the

formation of the active specie during the solvent-free transesterification. One study

proposed that the interaction of the solid sucrose with the solid catalyst leads to the
formation of solid sucrate, which further reacts with FAME through a solid-liquid process to

form sucrose monoester [14]. A more recent study suggested that the solid sucrose

dissolves at some extent in FAME, and that dissolved sucrose and liquid FAME can

interact on the catalyst surface where the reaction takes place [16]. In fact, according to

that study the solubility of sucrose in FAME is enhanced by temperature and the presence

of surfactants. Regardless the approach used to explain the solvent-free transesterification

mechanism, interfacial area and the proper stirring play a major role in the rate of reaction.

A poor agitation in the reactive media (low stirring velocities) can generate a larger mass

transfer boundary layer thickness, and therefore an external mass transfer limitation for

reaction. At the same time, a higher surface area of interaction enhances mass transfer

and increases reaction effectiveness. In this regard, the reaction must be operated at

sufficient high stirring velocities with small enough particles to ensure that the reaction is in

the kinetic regime [30]. As observed in Equation 4, the mass transfer coefficient (kc) in

solid-liquid media can be improved by increasing stirring velocity (v) or by decreasing solid

particle size (Dp).

)
$ (
! ~ #% ' (4)
&

Because of the use of the Ultraturrax® mixer, it was considered that the reacting media

was well homogenized before reaction, with small enough particles obtained by the

grinding effect of the mixer blades. However, preliminary reaction experiments operating at

different stirring rates allowed verifying that the kinetic profiles did not change above 500

rpm. For this reason, further kinetic experiments were performed at 700 rpm. Additional

experiments were performed to assure absence of mass transfer limitations associated


with sucrose particle size. A set of reactions were carried out with different sucrose particle

sizes at two different temperatures, and the results of these experiments are summarized

in Figure 3. The average reaction rate was measured in terms of the productivity of the

reaction. As observed, mass transfer limitations are negligible when using sucrose particle

sizes below 100 µm. Then, further kinetic experiments were performed using sucrose

particle sizes below this level.

Figure 3. Productivity of sucrose ester in the heterogeneous transesterification of sucrose

and methyl palmitate as a function of sucrose particle size, at different temperatures (

409K; 389K)

3.2. Reaction kinetics

According to the compiled results of Table 2, the average conversion obtained during

experiments was ~36%, which agrees with similar studies using surfactants (~ 40%) [13].

The highest productivity was 55 g SE/h kg which is below to the obtained when using

solvents as contacting agents (~100 g SE/h kg [6]). However, the maximum content of

sucrose ester on the final mixture was around 25%wt, which is similar to the obtained with

the solvent process. Despite the solvent-free process exhibits slightly lower productivity
than the solvent process, there is no need for solvent removal and the required

separations are less energy demanding.

Table 2. Operating conditions and results of kinetic experiments during sucrose esters

production using surfactants as contacting agent.


¥
T %wt Reaction %wt Monoester %wt SE Total FAME
Run Surfactant §
(K) Surfactant time (h) in total SE Final conversion (%)
1 5.02 4.43 51.69 15.28 36.86
2 409 9.41 3.90 67.64 18.70 49.16
3 12.73 3.90 66.11 21.26 39.75
4 4.46 23.53 40.71 13.41 52.78
5 KP 389 8.64 20.97 72.87 13.18 38.71
6 12.66 20.33 45.94 15.00 41.30
7 4.99 101.77 100.00 1.17 2.16
8 369 9.10 101.77 100.00 1.03 3.70
9 12.49 78.45 100.00 4.18 5.76
10 5.01 3.90 17.73 10.29 22.86
11 409 10.00 4.52 46.17 11.58 54.44
12 15.83 4.52 58.76 10.23 50.00
13 5.01 31.52 44.63 16.29 46.00
14 GMS 389 8.51 23.35 65.58 11.48 33.76
15 12.69 22.05 37.79 15.61 53.54
16 5.02 76.10 67.80 5.87 19.70
17 369 8.34 76.10 84.91 4.96 28.66
18 12.77 78.45 62.99 13.93 29.08
19 4.82 3.92 60.19 9.48 13.43
20 409 9.27 5.25 63.70 18.05 37.17
21 16.15 3.92 52.50 22.86 43.09
22 5.07 19.35 43.92 15.05 37.91
23 SE 389 9.16 20.87 59.92 22.20 45.30
24 12.47 20.87 44.10 25.16 40.12
25 4.98 71.58 37.20 13.77 32.13
26 369 9.27 101.77 49.47 14.39 61.31
27 13.13 71.58 71.32 9.15 57.58
28 KP 0%* 5.2 4.82 66.52 2.25 2.52
29 KP 1.1%* 409 5.07 4.60 45.37 15.49 39.62
30 KP 2.45%* 4.98 4.60 73.52 18.05 27.52
*Catalyst concentration
¥
Referred to the total weight of reacting mixture
§
With respect to methyl palmitate

The overall reaction of sucrose with FAME can be described with equation 5. As the

process was carried out under vacuum, methanol was continuously removed from the

reactive medium, thus avoiding the reversible reaction.


8997:1, <=>71??@A
*+,-./0(1 + 23456(7 BCCCCCCCCCCCCCCCD 5.2.0/0-(7 + EFGH.0/0-/(7 + 50ℎJ2.F($ (5)

As presented in Figure 1, the reaction occurs in a complex series-parallel scheme. For this

reason, the interpretation of experimental results requires some simplifications. In this

case, the kinetic model of sucrose ester production was constructed based on the

following assumptions:

1. The monoester production reaction is a solid/liquid reaction between solid sucrose

and liquid methyl palmitate. On the other hand, oligoester production is a liquid/liquid

reaction between liquid monoester (melting point of sucrose monoester is around 328-

338K [4]) and liquid methyl palmitate.

2. The heterogeneous characteristic of the production of monoester was taken into

account by mean of an additional reaction step (sucrose activation), similarly to a

previous study [14]. In the literature, this activation step was used as a model of a

solid-solid interaction between the catalyst and sucrose. In this work, this additional

step represents the dissolution enhancement by surfactant addition, as well as the

encounter between the dissolved sucrose and the solid catalyst. Both approaches

result in a similar power law kinetic expression. However, here, the catalyst and

surfactant concentrations are taken into account within the reaction rate expression.

This is an improvement in the kinetic model because the effect of catalyst and contact

agent concentrations on the reaction rate can be predicted.

3. The saponification of esters with K2CO3 was assumed negligible because this catalyst

is a non-saponifying alkali [6]. If there is no water on the system, saponification does

not occur.
4. Sucrose and/or sucrose ester oligomerization, and hydrolysis are considered

negligible.

5. Though the transesterification reaction is reversible, it was considered irreversible

under the operating conditions of kinetics experiments. Continuous vapor removal

caused by the vacuum operation ensured a negligible methanol concentration in the

reactive media.

6. Since the oligoesters formation occurs as liquid-liquid reactions, their reaction rates

were assumed similar. Consequently, polysubstituted sucrose esters were treated as

a single component regardless the degree of substitution. This assumption was used

in other kinetic studies on sucrose ester production [6, 14]. The size of this oligoester

was fixed with parameter 2. This was set depending on the experimentally obtained

substitution degrees in the esters mixture (excluding monoester). This parameter

corresponds to the number of FAME moles needed to obtain an oligoester with an

average substitution degree of 2 + 1, as described in the stoichiometric coefficient of

FAME in Equation 8 of Table 3.

7. Mass transfer limitations were negligible (i.e. kinetic regime).

8. The reaction rates were described using an elementary power law kinetic model, so

that the reaction order corresponded to the stoichiometric coefficients.

9. Surfactants do not react at the conditions of the experiments. Theoretically, hydroxyl

groups of glycerol monoesters could react with FAME to form glycerol diesters.

However, concentration of glycerol monostearate was quantified and no considerable

changes along reaction time were observed (See Figure S-3 of Supplementary

Material).

Based upon these assumptions, the reaction model describing sucroesters synthesis is

summarized in Equations 6 to 8 of Table 3. In these equations subindex j represents the


reaction step, rj is the rate of reaction (1/s), and kj is the rate constant (1/s). The mole

fractions in the model corresponds to sucrose (SuOH, x1), activated sucrose (SuOH*, x2),

methyl palmitate (RCOOMe, x3), sucrose monoester (SuO(COR), x4), methanol (MeOH, x5

= 0), catalyst (K2CO3, x6), surfactant (Emu, x7), and sucrose oligoesters (SuO(COR)n, x8).

Table 3. Reactions steps and kinetic model of the transesterification of sucrose and methyl

palmitate using different surfactants as contact agents.

j Reaction Reaction rate

1
Sucrose
SuOH + K2CO3 + Emu  SuOH* + K2CO3 + Emu - = (K L   (6
activation
-( = ( (  (7
Monoester
2 SuOH*+ RCOOMe  SuO(COR) + MeOH
production
- =   (8
Oligoester O
3 SuO(COR) + n RCOOMe  SuO(COR)n+1 + n MeOH
production

In order to verify the dependence of the rate of reaction with the catalyst and surfactant

concentration, some preliminary experiments were carried out varying their concentrations.

In each preliminary experiment, the rate of sucrose disappearance at the corresponding

reaction time required to achieve 50% of final conversion was calculated. This value was

used instead of the initial rate because during the lag period of activation the conversion

was negligible and therefore not comparable. Sucrose concentration with time was

calculated with Equation 1, and the disappearance rate was calculated as the slope of the

fitted equation to the experimental sucrose concentrations (See Table S-1 in

Supplementary Material). Results obtained at 409K are presented in Figure 4 and for other

temperatures, in Figures S-4 and S-5 in the supplementary material.

While the expected linear dependence with the catalyst loading was observed for

concentrations below 3%wt, some deviations were observed at higher loadings (See

Figure 4a). This nontypical behavior was also observed in one of the earlier studies on the

production of sucrose esters [28]. In that study, the behavior of the reaction using solvent
at higher concentrations of potassium sucrate (active species formed from the contact of

the base and sucrose) was non-ideal because the rate of reaction was no longer directly

proportional to the sucrate ion concentration. In the reported study it was claimed that the

solid potassium sucrate precipitated at high concentrations as a results of the association

of the ionic species, thus reducing the effective concentration of the catalyst. In the case of

the solvent-free reaction, a similar reason could explain the reduction of the rate of

reaction at high catalyst loadings. Also, it is expected that high concentrations of catalyst

could limit the proper contact between reactants due to the increase in the solid fraction in

the reaction media. In the present work, it was observed a loss in the homogeneity in the

reactive media at higher catalyst concentrations. After the homogenization in the

Ultraturrax®, the catalyst remained dispersed in most experiments. However, under high

catalyst loadings, the solid particles tended to form agglomerates and to settle, thus

reducing their interaction with reactants. Kinetic parameters regressed in this work

correspond to experiments performed with 5 %wt catalyst. Therefore, the non-linear

behavior was not considered in the model. This indicates that the use of the proposed

kinetic model at different catalyst concentrations other than 5%wt has to be carefully

considered.

Regarding the surfactant concentration, a linear dependency of the reaction rate on

concentrations was verified (See Figure 4b). The proportionality was considerably higher

for potassium palmitate than for the other emulsifiers. This was mainly caused by the

catalytic effect of this soap, which will be further discussed. The effect of surfactant

concentration was diminished at low temperatures (not in the Figure 4b).

The dependence of the reaction rate constants with temperature was calculated with an

Arrhenius-type expression as shown in Equation 9.


RST,U

Q = 4Q 0 VW (9)

Herein, R is the ideal gas constant, T is the absolute temperature, and Aj and Ea,j are the

pre-exponential factor and the activation energy of the reaction, respectively.

Consequently, each reaction has two kinetic parameters to be determined in the

parameter regression. The rate of reaction for oligoester production has the additional

parameter 2, corresponding to the moles of methyl ester consumed. However, this

parameter was not included in the parameter regression because it was determined from

the average esterification degree of the final product for each emulsifier used as contact

agent.

3.3. Kinetic modeling

The kinetic parameters were estimated by fitting the experimental data (composition of

chemical species in mole fraction) with the numerical integration of model in a batch

reactor, expressed as follows:

XYZ XZ
=[ = ∑Y^
Q_ ]?,Q -$,?,Q (10)
X X

where [? is the number of moles of component i, [ is the total number of moles in the

reactor, ? is the mole fraction of component i, ]?,Q is the ratio of stoichiometric coefficients

of component i, and -$,?,Q is the volumetric reaction rate of component i in reaction j. Due to

mass excess of methyl palmitate, the mixture density can be considered constant. In

addition, in these reactions the total number of moles is conservative, and then Equation

10 can be simplified to Equation 11.


XZ
= ∑Y^
Q_ ]?,Q -$,?,Q (11)
X

The corresponding mole balances for the components involved in reaction are showed in

equations 12 to 19.

Figure 4. Rate of sucrose disappearance in sucrose ester production at 409K using a)

5%wt potassium palmitate as a function of catalyst’s concentration and b) 5%wt catalyst

as a function of surfactant’s concentration. ( potassium palmitate; glycerol

monostearate; sucrose ester)


X`
X
= −- (12)

Xa
= - (13)
X

X
X
= −-( − 2 - (14)

X
X
= -( − - (15)

X
X
= - (16)

Xb
X
= - − - = 0 (17)

Xc
= - − - = 0 (18)
X

Xd
X
= -( + 2 - (19)

The system of differential equations was integrated using Matlab®, with an ode45 function

(fourth-order Runge-Kutta). The initial mole fraction of all components was used as the

boundary condition to integrate the first order differential equations. The kinetic parameters

were determined by the minimization of the objective function defined in Equation 20. This

corresponds to the sum of squares of the difference between the experimental and

calculated molar fractions of all the components of the reactive media.


(
Z,ef& Z,gThg 
3=?O = ∑Y8
?_ )[j (20)
iZ

The experimentally measured concentrations corresponded to methyl palmitate, sucrose

monopalmitate and polysubstituted sucrose esters. Concentrations of polysubstituted

sucroesters were lumped to obtain the concentration of the here named sucrose

oligoester. As the order of magnitude of the concentrations of all components changed by

a factor of 10, the parameter 3? was used in the minimization function to normalize the

concentration values. Using this approach, the error of the model was equally minimized

for all three components. Their values were fixed as 3 = 1, 3( = 0.1 and 3 = 0.01. The

function was minimized using the fminsearch function included in Matlab® based upon the

Nelder-Mead simplex algorithm. A preliminary regression was carried out in Microsoft

Excel® where the Arrhenius behavior of the rate constants was verified, as shown in Figure

5. From this plot, initial guess values for the energy of activation and the pre-exponential

factor were obtained, and used as seed values during minimum squares regression.

The final set of kinetic parameters are presented in Table 4. The values of the activation

energies are higher than those reported for alkaline transesterification using the solvent

process (41421 J/mol, [28]), and slightly lower than those observed in the enzymatic

transesterification using solvents (203360 J/mol, [20]). According to the values of 2 used in

this work, oligoesters obtained had an average esterification degree (2 + 1) of 3.2. The

value of this parameter was not considerably affected by type of surfactant, so oligoesters

mostly corresponded to a mixture of sucrose tri- and tetrapalmitate. For this reason, the

value of 2 was kept fixed for all experiments with the same contact agent.
Figure 6 are parity plots comparing the whole set of experimental data with calculations

from the developed models. As observed, there is good agreement between the obtained

kinetic expressions and the experimental observations. To verify the good agreement of

the models, in Figure 7 it is presented a comparison between experimental and adjusted

kinetic profiles at 409K using 15 %wt of the different surfactants. The complete set of

kinetic plots for all the experiments described in Table 2 are included in Figure S-6 to S-32

in the supplementary material.

Table 4. Kinetic parameters for heterogeneous transesterification of sucrose and methyl

palmitate with different surfactants.

Potassium Confidence Glycerol Confidence Sucrose Confidence


Parameter Units
palmitate interval monostearate interval ester interval
4 1/s 7.606 x 1011 ±1.437 x 1010
2.103 x 108
±4.168 x 106 4.810 x 108 ±1.052 x 107
4( 1/s 6.991 x 1017 ±1.321 x 1016 4.309 x 1014 ±8.539 x 1012 1.108 x 108 ±2.424 x 106
4 1/s 1.576 x 1013 ±2.977 x 106 2.199 x 1014 ±4.358 x 1012 154 923 ±3389
69, J/mol 109 944 ±2078 86 276 ±1710 81 619 ±1785
69,( J/mol 166 448 ±3145 125 315 ±2483 95 534 ±2090
69, J/mol 93 457 ±1766 139 353 ±2762 73 624 ±1610
2 2.4 2.1 2.3
3=?O 0.1004 0.155 0.3132
Average Error % 3.4 3.5 3.0
Number of
83 79 98
Experimental data

In the kinetic profiles, it is observed that there is an induction time (lag period) at the

beginning of the reaction when the concentrations of reactants barely change, followed by

rapid products formation. This is typical of some heterogeneous reactions [6, 11], and it

was indirectly considered in the reaction mechanism through the sucrose activation that

was the limiting step. According to results, the extent of this induction time was heavily

influenced by temperature, and slightly by catalyst and surfactant concentration.


In the specific case of the kinetic profiles obtained when using glycerol monoester as

surfactant, large deviations between experimental data and model prediction were

observed under high surfactant loadings. In this case, the suspension of solids in the liquid

phase was not stable during the whole reaction time. With the advance of the reaction,

agglomeration of solids was observed, and in some cases (i.e. with high surfactant

content) this effect was so strong, that the reactive media became heterogeneous, and the

solids could not be dispersed any longer.

Despite all the models exhibited reasonably good agreement with experiments, the kinetic

model obtained for sucrose ester as surfactant showed the larger deviations. This can be

explained by the reactivity of the surfactant. The composition profiles indicated that the

loaded sucrose monoester (when sucrose ester added was more than 5 %wt) disappeared

as soon as the reaction started. This can be caused either by its reaction into oligoesters

(considered in the model) or by its thermal decomposition. Since oligoesters were not

observed, decomposition (e. g. thermal degradation, hydrolysis) is a more probable cause

for this rapid disappearance of sucrose monoester. Despite this side reaction, the

production of sucrose monoester is faster after the lag period and the decomposition

seems to be compensated. With the aim of a future improvement in the kinetic model, the

side reactions should be taken into account when sucrose ester is used as surfactant.
a

c
Figure 5. Arrhenius plots of a) sucrose activation, b) monoester production and c)

oligoester production, using 5%wt catalyst and different surfactants ( potassium

palmitate; glycerol monostearate; sucrose ester)

c
Figure 6. Parity plots of experimental and calculated concentrations in the

transesterification of sucrose and methyl palmitate, using different surfactants: a)

Potassium palmitate b) Glycerol monostearate. c) Sucrose ester, ( methyl palmitate;

sucrose monopalmitate; sucrose palmitate oligoesters)

Figure 7. Kinetic profiles of the transesterification of sucrose and methyl palmitate at 409K

using 15%wt surfactant: a) Potassium palmitate (run 3) b) Glycerol monostearate (run 12)
c) Sucrose ester (run 21). ( methyl palmitate; sucrose monopalmitate; sucrose

palmitate oligoesters)

3.4. Effect of Temperature

A strong effect of temperature on the solvent-free transesterification of sucrose and methyl

palmitate was experimentally observed. In fact, reaction times to obtain sucrose ester near

20 %wt in the reactor changed from 4h at 140°C to 20h at 120°C, approximately.

According to the Arrhenius expression, the larger the activation energy, the more

temperature-sensitive is the rate of reaction [30]. Considering the regressed parameters

summarized in Table 4, all reactions exhibited a strong temperature dependence.

Particularly, the temperature dependence of sucrose activation can be due to the

enhanced sucrose solubility at higher temperatures. Previous researches showed that

sucrose solubilization in methyl laurate increased by a factor of 10 when temperature

change from 343 to 403K in a system using 2.5%wt of potassium stearate as surfactant

[16]. This effect is also observed in Figure 8, where productivities of reactions are plotted

as a function of the reaction temperature. Productivities are improved in the same factor as

reaction rates when temperature is increased. For further process design, it is

recommended to carry out the reaction at temperatures near 409K.

3.5. Effect of type of surfactant and concentration

When comparing performance of the different surfactants in terms of productivity during

reactions (See Figure 8), potassium palmitate exhibited the best effectiveness. This can be

explained by the reported synergetic catalytic effect of potassium palmitate in reaction [13,

31]. This was verified by running the experiment without catalyst (See Table 2), where

production of sucrose ester was evidenced. This was also observed in studying the linear
dependence of reaction rate with catalyst concentration (See Figure 4a), where even in

absence of catalyst, there was some sucrose conversion. The proposed kinetic model was

used only for experiments with 5 %wt catalyst. Therefore, the catalytic activity of

potassium palmitate without catalyst can’t be described through this model. For further

process design, the catalytic effect of potassium palmitate should be studied in order to

assess their synergistic effects with the K2CO3 catalyst.

According to the experimental observations, the reaction rate followed a linear

dependence with the contact agent concentration (See Figure 4b). This dependence is the

highest for potassium palmitate among the studied emulsifiers. This is evidenced in the

value of the pre-exponential factor for the activation reaction (4 ) which was the highest for

potassium palmitate (See Table 4). Therefore, besides the frequency of collisions, the

higher pre-exponential factor for this emulsifier accounts for the intrinsic catalytic effect

and the contacting enhancement achieved with the surfactant.

Despite the high productivity obtained when using potassium palmitate as contact agent,

there are some challenges to overcome. At high concentrations, this soap might generate

undesirable flavors when used in food or feed products. Then, if the potassium palmitate is

not removed from the sucroesters, the application of the final product might be restricted

for cleaning or cosmetics products. This separation might involve energy intensive and

materials demanding processes losing the advantages over the solvent Process. Then if

required for food applications, the use of other food-compatible alkaline or earth alkaline

soaps as contacting agents (e. g. sodium and magnesium stearates) could be explored

using the methodology here described.


a

Figure 8. Productivity of the transesterification of sucrose and methyl palmitate at 409K

using 10%wt surfactant: as a function of temperature and surfactant’s concentration. a)


Potassium palmitate b) Glycerol monostearate. c) Sucrose ester, ( 369K, 389K,

409K). From results reported in Table 2

According to the results, the average monoester content in the final products was 62%wt

for potassium palmitate, 54%wt for sucrose ester and 41%wt for glycerol monostearate

(calculated from values in Table 5). Then, the calculated HLB values of the final product of

each surfactant (approximated from [32]) were 12 for reactions with potassium palmitate,

10 when using sucroester as surfactant, and 7 for glycerol monostearate. This indicated

that the surfactant has a strong effect in the selectivity of reaction, and in the degree of

substitution of final product.

As aforementioned, the effect of emulsifier in the selectivity of the reaction was not

evidenced in the value of the parameter 2. However, some analysis can be done regarding

the other kinetic parameters. In accordance with the model, the selectivity of monoester

with respect to oligoesters is defined by equation 21.

X ⁄X
*l/n = X⁄X (21)


Using the kinetic parameters obtained with the regression of the experimental data, the

selectivity to monoester along reaction was calculated at 409K and 5%wt of surfactant,

and the results are reported in Figure 9.


Figure 9. Predicted effect of type of surfactant on the selectivity to monoester in reactions

at 409 K and 5%wt of surfactant (―) Potassium palmitate. (―) Glycerol monostearate.

(―) Sucrose ester)

As observed, average selectivity at the corresponding reaction time required to achieve

50% of final conversion (c.a. 2h) agrees with the experimental observations. Results

confirm that the product with a higher content of sucrose monoester is produced when

using potassium palmitate as surfactant. On the other hand, sucrose esters with higher

degree of esterification were obtained when using glycerol monostearate. This can be

explained by the microenvironment of the produced emulsions. Potassium palmitate has

the highest HLB (~ 20), so that it enhances the interaction of the more polar molecules (i.e.

sucrose) with the continuous phase (i.e. methyl ester). In contrast, glycerol monostearate

has a low HLB (~ 4), so that it promotes transesterification of more hydrophobic molecules

within the reactive media (i.e. oligoesters). Despite the different kinetic behaviors, the

obtained models are able to predict productivity and selectivity under the different

operating conditions, and they can be preliminary used for further process design and up-

scaling.

4. Conclusions

The kinetics of heterogeneous transesterification of sucrose and methyl palmitate to

produce sucroesters was evaluated. Potassium palmitate, sucrose palmitate or glyceryl

monoester were used to enhance compatibility between reactants during reaction. The

kinetic effects of temperature (369 to 409 K) and surfactant’s type and concentration (5 to
15%wt) were assessed. Higher productivities of sucroesters and higher selectivity to

monoesters were obtained when potassium palmitate was used as surfactant. This

component also showed a catalytic activity enhancing reaction performance. The lower

monoester content in the final product was obtained when using glycerol monostearate as

surfactant. The reaction model was described using a sucrose activation step, which was

consistent with the observed lag period at the beginning of reaction. Kinetic models were

obtained for the transesterification reaction with the different surfactants, and they showed

good agreement with experiments, especially for potassium palmitate. These models can

be used for further analysis and preliminary design of a solvent-free process in the

production of biobased sucrose esters.

Nomenclature

List of symbols and Abbreviations

4 Arrhenius frequency factor of sucrose activation 1⁄/


4( Arrhenius frequency factor of monoester production 1⁄/
4 Arrhenius frequency factor of oligoester production 1⁄/
pq Particle diameter ,r
69, Activation energy of sucrose activation s/r.F
69,( Activation energy of monoester production s/r.F
69, Activation energy of oligoester production s/r.F
[? Number of moles of component i r.F .t G
! Mass transfer coefficient ,r⁄/
 Rate constant of sucrose activation 1⁄/
( Rate constant of monoester production 1⁄/
Rate constant of oligoester production 1⁄/
- Rate of reaction sucrose activation 1⁄/
-( Rate of reaction monoester production 1⁄/
- Rate of reaction oligoester production 1⁄/
-?,Q Reaction rate of component i in reaction j 1⁄/
-$,?,Q Volumetric reaction rate of component i in reaction j 1/,r ∙ /
*l/n Instantaneous selectivity of monoester with respect to
oligoester
v Velocity of the reactive mixture ,r⁄/
? Molar fraction of component i
ACN Acetonitrile
DMSO Dimethylsulfoxide
HPLC High Performance Liquid Chromatography
Emu Surfactant Subindex 7
GMS Glycerol monostearate
K2CO3 Potassium carbonate (catalyst) Subindex 6
KP Potassium palmitate
MeOH Methanol Subindex 8
RCOOMe Methyl palmitate Subindex 3
SE Sucrose ester
SuO(COR) Fatty acid sucrose monoester Subindex 4
SuO(COR)n Fatty acid sucrose oligoester Subindex 5
SuOH Sucrose Subindex 1
SUOH* Active sucrose Subindex 2
[ Total number of moles in the reaction mixture r.F
[j Number of components measured
[w Number of reactions
3? Normalization factor
3=?O Minimization function
w Ideal gas constant s/r.F ∙ x
y Reaction temperature x
z Volume ,r
 Time /

Greek letters

]?,{ Ratio of stoichiometric coefficients of component i in reaction k

Subscripts and Superscripts

G Component identification 1 to 8

| Reaction identification 1 to 3

,JF, Calculated data

0} Experimental data

Acknowledgement
This work was financially supported by Colciencias, through the program Convocatoria

617 de 2014 para apoyo a proyectos de doctorado en Colombia and through the program

Jóvenes Investigadores Convocatoria 706 de 2015.

References

[1] R.J. Farn, Chemistry and Technology of Surfactants, Backwell Publishing Ltd, 2006.

[2] I.J. a. Baker, R.I. Willing, D.N. Furlong, F. Grieser, C.J. Drummond, Sugar fatty acid

ester surfactants: Biodegradation pathways, J. Surfactants Deterg. 3 (2000) 13–27.

doi:10.1007/s11743-000-0108-1.

[3] D. Reyes-Duarte, N. López-Cortés, M. Ferrer, F.J. Plou, A. Ballesteros, Parameters

affecting productivity in the lipase-catalysed synthesis of sucrose palmitate,

Biocatal. Biotransformation. 23 (2005) 19–27. doi:10.1080/10242420500071763.

[4] N. Otomo, Basic properties of sucrose fatty acid esters and their applications, in:

D.G. Hayes, D. Kitamoto, D.K.Y. Solaiman, R.D. Ashby (Eds.), Biobased

Surfactants Deterg., AOCS Press, Urbana, Illinois, 2009: pp. 275–298.

[5] Y. Shi, J. Li, Y.-H. Chu, Enzyme-catalyzed regioselective synthesis of sucrose-

based esters, J. Chem. Technol. Biotechnol. 86 (2011) 1457–1468.

doi:10.1002/jctb.2711.

[6] P.S. Deshpande, T.D. Deshpande, R.D. Kulkarni, P.P. Mahulikar, Synthesis of

Sucrose–Coconut Fatty Acids Esters: Reaction Kinetics and Rheological Analysis,

Ind. Eng. Chem. Res. 52 (2013) 15024–15033. doi:10.1021/ie401524g.

[7] H.B. Hass, N.J. Summit, F.D. Snell, W.C. York, L.I. Osipow, Process for producing

Sugar Esters, 2893990, 1959.

[8] W.A. Farone, R. Serfass, S. Robert, Sugar-ester manufacturing process, US

5756716, 1996.

[9] K.S. Nakamura, S.H. Nagahara, J.O. Kawaguchi, Process for production of high-
monoester sucrose higher fatty acid esters, 5453498, 1995.

[10] L.I. Osipow, W. Rosenblatt, S. Valley, Esterification of polyhydric compounds in the

presence of transparent emulsifying agent, 3480616, 1969.

[11] J. Fitremann, Y. Queneau, J. Maitre, A. Bouchu, Co-melting of solid sucrose and

multivalent cation soaps for solvent-free synthesis of sucrose esters, Tetrahedron

Lett. 48 (2007) 4111–4114. doi:10.1016/j.tetlet.2007.04.015.

[12] K. James, J.F. Smith, Process for the production of esters of sugar and sugar

derivatives, 2009.

[13] K. James, J.F. Smith, Process for the production of esters of sugars and sugar

derivatives, 8329894 B2, 2012.

[14] A.L. Le-Coent, M. Tayakout-fayolle, F. Couenne, S. Briancon, J. Lieto, J. Fitremann-

Gagnaire, Y. Queneau, A. Bouchu, Kinetic parameter estimation and modelling of

sucrose esters synthesis without solvent, Chem. Eng. Sci. 58 (2003) 367–376.

[15] H.R. Galleymore, K. James, H.F. Jones, C.L. Bhardwaj, Process for the production

of a surfactant containing sucrose esters, 4298730, 1981.

[16] R. Zhao, Z. Chang, Q. Jin, W. Li, B. Dong, X. Miao, Heterogeneous base catalytic

transesterification synthesis of sucrose ester and parallel reaction control, Int. J.

Food Sci. Technol. 49 (2014) 854–860. doi:10.1111/ijfs.12376.

[17] N. Desai, B. Grüning, Process for the preparation of sucrose fatty acid esters,

5945519, 1999.

[18] J. Fitremann, A. Bouchu, Y. Queneau, Synthesis and Gelling Properties of N-

Palmitoyl-L-Phenylalanine Sucrose Esters, Langmuir. 19 (2003) 9981–9983.

[19] K.J. Parker, K. James, J. Hurford, Sucrose Ester Surfactants — A Solventless

Process and the Products Thereof, in: J.L. Hickson (Ed.), Sucrochemistry, American

Chemical Society, Washington, DC, 1977: pp. 97–114.

[20] C.F. Braga, C.A.M. Abreau, N.M. Lima-Filho, Y. Queneau, G. Descotes, Kinetics of
Catalytic Sucrose Esterification in Aqueous Media, React. Kinet. Catal. Lett. 77

(2002) 91–102.

[21] S. M., K. Tatsumi, Effect of Sucrose Ester of Fatty Acid on the Antioxidant Activity of

Milk Products on Fish Oil Oxidation, J. Food Sci. 67 (2002) 547–552.

[22] J. Aburto, I. Alric, E. Borredon, Organic Solvent-free Transesterification of Various

Starches with Lauric Acid Methyl Ester and Triacyl Glycerides, Starch - Stärke. 57

(2005) 145–152. doi:10.1002/star.200400380.

[23] F.J. Plou, M.A. Cruces, M. Ferrer, G. Fuentes, E. Pastor, M. Bernabé, M.

Christensen, D. De Biocatálisis, I. De Catálisis, I.D.Q. Orgánica, A.S. Novozymes,

N. Allé, I. De Investigaciones, A. De Barcelona, D. Catálisis, C.U.A.M. Cantoblanco,

Enzymatic acylation of di- and trisaccharides with fatty acids: choosing the

appropriate enzyme, support and solvent, J. Biotechnol. 96 (2002) 55–66.

[24] S. Sugar-Research-Foundation, Sucrose Esters Report, 1961.

[25] J. Van Gerpen, Biodiesel processing and production, Fuel Process. Technol. 86

(2005) 1097–1107. doi:10.1016/j.fuproc.2004.11.005.

[26] P.C. Narváez, S.M. Rincón, F.J. Sánchez, Kinetics of palm oil methanolysis, J. Am.

Oil Chem. Soc. 84 (2007) 971–977. doi:10.1007/s11746-007-1120-y.

[27] Q. Wang, S. Zhang, P. Zhang, J. Zhu, J. Yang, Separation and Quantitation of

Sucrose Esters Using HPLC with Evaporative Light Scattering Detection, J. Liq.

Chromatogr. Relat. Technol. 29 (2006) 2399–2412.

doi:10.1080/10826070600864874.

[28] R.U. Lemieux, A.G. McInnes, The preparation of sucrose monoesters, Can. J.

Chem. 10 (1962) 2376–2393.

[29] U. Schuchardt, R. Sercheli, R.M. Vargas, Transesterification of Vegetable Oils : a

Review, J. Braz. Chem. Soc. 9 (1998) 199–210.

[30] H.S. Fogler, Elements of Chemical Reaction Engineering, 4th. ed., Pearson
Education Inc., 2006.

[31] R.O. Feuge, H.J. Zeringue, T.J. Weiss, Process for the production of sucrose esters

of fatty acids, 3714144, 1973.

[32] B.A.P. Nelen, J.M. Cooper, Sucrose esters, in: R.J. Whitehurst (Ed.), Emuls. Food

Technol., Backwell Publishing Ltd, 2004: pp. 131–162.

Figure Captions

Figure 1. Sucrose fatty acid esters production by transesterification of fatty acid methyl

ester (FAME) and sucrose

Figure 2. Experimental set-up kinetic experiments

Figure 3. Productivity of sucrose ester in the heterogeneous transesterification of sucrose

and methyl palmitate as a function of sucrose particle size, at different temperatures (

409K; 389K)

Figure 4. Rate of sucrose disappearance in sucrose ester production at 409K using a)

5%wt potassium palmitate as a function of catalyst’s concentration and b) 5%wt catalyst

as a function of surfactant’s concentration. ( potassium palmitate; glycerol

monostearate; sucrose ester)


Figure 5. Arrhenius plots of a) sucrose activation, b) monoester production and c)

oligoester production, using 5%wt catalyst and different surfactants ( potassium

palmitate; glycerol monostearate; sucrose ester)

Figure 6. Parity plots of experimental and calculated concentrations in the

transesterification of sucrose and methyl palmitate, using different surfactants: a)

Potassium palmitate b) Glycerol monostearate. c) Sucrose ester, ( methyl palmitate;

sucrose monopalmitate; sucrose palmitate oligoesters)

Figure 7. Kinetic profiles of the transesterification of sucrose and methyl palmitate at 409K

using 15%wt surfactant: a) Potassium palmitate (run 3) b) Glycerol monostearate (run 12)

c) Sucrose ester (run 21). ( methyl palmitate; sucrose monopalmitate; sucrose

palmitate oligoesters)

Figure 8. Productivity of the transesterification of sucrose and methyl palmitate at 409K

using 10%wt surfactant: as a function of temperature and surfactant’s concentration. a)

Potassium palmitate b) Glycerol monostearate. c) Sucrose ester, ( 369K, 389K,

409K). From results reported in Table 2

Figure 9. Predicted effect of type of surfactant on the selectivity to monoester in reactions

at 409K and 5%wt of surfactant (―) Potassium palmitate. (―) Glycerol monostearate. (―)

Sucrose ester)
Highlights

• Kinetics of solvent-free transesterification of sucrose with FAME was studied


• Three different surfactants were studied as contacting agents of reaction
• Three kinetic models were obtained for the different surfactants evaluated
• Reaction rate is highly temperature dependent.
• There is a relationship between HLB of contacting agent and HLB of final product

Você também pode gostar