Anda di halaman 1dari 10

Chemical Engineering Journal 169 (2011) 319328

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Kinetic modelling of the esterication of rosin and glycerol: Application to


industrial operation
Miguel Ladero a, , Miguel de Gracia a , Juan Jose Tamayo a , Inmaculada Lopez de Ahumada b ,
Fernando Trujillo b , Felix Garcia-Ochoa a
a
b

Dpt. Ingeniera Qumica. Fac. CC. Qumicas, Universidad Complutense, 28040-Madrid, Spain
La Unin Resinera Espa
nola S.A., 28001-Madrid, Spain

a r t i c l e

i n f o

Article history:
Received 9 December 2009
Received in revised form 20 January 2011
Accepted 7 March 2011
Keywords:
Rosin
Glycerol
Esterication
Kinetic model
Gel permeation chromatography
Industrial scale

a b s t r a c t
The esterication of glycerol and rosin is a reaction of importance in the coating and adhesive industries,
main consumers of the triglyceride of rosin acids. While this chemical transformation has led to some of
the most useful products of rosin, kinetics of the reactions involved have not been studied in depth. In this
study, a new analytical method based on gel permeation chromatography (GPC) is applied and compared
to classical acid titration, the industrial standard. Experimental data on rosin esterication with glycerol
were obtained at laboratory scale, covering a wide range of operational conditions, changing rosin to
glycerol molar ratio (between 2 and 4) and temperature (between 240 and 280 C). Afterwards, different
kinetic models were tted to experimental data to select between rst- and second-order potential and
hyperbolic kinetic models. Linear and non-linear regression techniques with numerical integration of
the differential equations were used to t the proposed kinetic models. A hyperbolic model coupled with
glycerol mass balance that considered stripping of this polyalcohol proportional to the global reaction
rate and to glycerol concentration proved to be the most adequate both at a given temperature and in
all the temperature intervals studied. On the other hand, samples were analysed by H NMR and ionic
chromatography to determine the glycerol amount in the reaction medium and in the distillate, respectively, observing the stripping of the polyalcohol. Finally, the selected model was used to simulate and t
runs executed at the usual industrial mode, a transient temperature operation followed by an isothermal
operation with gradual addition of glycerol (fed-batch operation). These runs were performed at lab and
at industrial scale, and the selected kinetic model and glycerol mass balance were adequate to simulate
rosin conversion.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Rosin or colophony is the non-volatile part of the resin of pines
and some other conifers. It is mainly composed (70%) of abietic
acid, pimaric acid and their isomers (as well as some hydrogenated
and dehydrogenated acids of the same family). Some neutral compounds (alcohols, aldehydes, esters) are also present [1]. There are
several kinds of rosin. The most pure comes from the secretion of
pine trees that are bled to obtain resin, subject to a further distillation process at 170180 C. The heavier fraction of such process is
gum rosin. A similar process beginning with chipping the trunk of
pines when they are cut down leads to the wood rosin. Purication
processes of the black liquor in the Kraft paper manufacture give
rise to the so-called tall-oil rosin. All of them have qualitative similarities in their composition, but the highest concentration of rosin

Corresponding author. Tel.: +34 913944164; fax: +34 913944179.


E-mail address: mladero@quim.ucm.es (M. Ladero).
1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.03.012

acids is found in gum rosin (70% rosin acids and 30% terpenic
hydrocarbons), whose main world producer is the Peoples Republic of China (approx. 90% world production: 450,000 ton/year gum
rosin) [2].
Rosin (and most of its products) is used in the paper, coating
(varnishes, wax, and adhesives), polymer and food industries, as
well as a precursor for ux in soldering [36]. With bases, rosinates
are obtained, being extensively used as soaps. With dicarboxilic
acids (maleic and fumaric acids, mainly), some adducts are produced, to gain stability towards oxidation [1,7]. Dimerization of
rosin acids is another strategy to reduce the number of double
bonds and to enhance oxidative stability [1,8]. By using hydrogen
and medium pressures, or simple heating with or without some
catalysts, hydrogenated and disproportionated rosin are manufactured (for the polymer industry) [1,9]. Esterication with methanol
yields ethylene-glycol, glycerol, and pentaerythritol, esters for a
variety of applications [1,10]. These are used as tackiers for hotmelt and pressure adhesives, in solder uxes, as crystallization
promoters in the production of polypropylene, as neutralizers n

320

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

Nomenclature
glycerol area in H-RMN spectra
ester area in H-RMN spectra
colophony or rosin area in gel permeation chromatographs
AIC
Akaikes information criterion
AICc
Akaikes information criterion for small set of data
%A
percentual area of peaks in gel permeation chromatographs
1 , 2 , t0 parameters in Boltzmann equation for the tting
of temperature with time
1 , 2 , 3 parameters in a modied hyperbola for the tting
of temperature with time
colophony or rosin concentration (mol L1 )
CC
CG
glycerol concentration (mol L1 )
CG0
initial glycerol concentration (mol L1 )
13 C NMR carbon-nuclear magnetic resonance spectroscopy

chemical shift in NMR (ppm)


activation energy (J mol1 )
Ea
F95
Fischers F value at 95% condence
FG
glycerol addition ow in transient temperature runs
(mol min1 )
GPC
gel permeation chromatography
H NMR hydrogen-nuclear magnetic resonance spectroscopy
HPLC
high performance liquid chromatography
k
rst-order kinetic constant (min1 )
k 0
pre-exponential term of the rst-order kinetic constant (min1 )
k
second-order kinetic constant (L mol1 min1 )

pre-exponential term of the second-order kinetic
k 0
constant (L mol1 min1 )
second-order kinetic constant of the hyperbolic
k1
model (L mol1 min1 )
pre-exponential term of the second-order kinetic
k10
constant of the hyperbolic model (L mol1 min1 )
coefcient for the stripping of glycerol in glycerol
kst
mass balances 2 and 3
K2
denominator constant of the hyperbolic model
(L mol1 )
K
number of parameters in a given model
M
molar concentration ratio between glycerol and
rosin
N
number of experimental data
nC
rosin amount of substance (mol)
nC0
rosin initial amount of substance in transient temperature runs and in Eq. (8) (mol)
SQR
sum of squares of weighted errors
SSQWPV sum of squares of weighted predicted values
PID
proportional-integral-differential controller
r
reaction rate (mol L1 min1 )
disappearance rate of rosin (mol L1 min1 )
RC
RG
disappearance rate of glycerol (mol L1 min1 )
T
temperature ( C)
t
time (min)
initial temperature in a transient temperature run
T0
( C)
V
reaction mixture volume (L)
rosin conversion
XC
XC 240
rosin conversion at time = 240 min.
XG 240
glycerol conversion at time = 240 min.
glycerol conversion
XG
AG
AE
AC

the paper industry, in the formulation of chewing gum, and for


the manufacture of polymeric coatings used in the controlled dispersion of drugs and tosanitary products, and some other minor
applications [3,6,11,12].
Glycerol is a highly hygroscopic polyalcohol (1,2,3propanetriol), with a high viscosity and relatively high density, and
with several applications in the cosmetic, food, pharmaceutical,
and chemical industries (polymers, triacetin, and more). However,
it is used in relatively small amounts in every application so
its production has remained relatively low (600,000 ton/year).
In the last years, glycerol has been produced as a by-product
in the biodiesel industry, resulting in a signicant increase in
its production and the build-up of large stocks of this chemical.
Consequently, the price of this polyalcohol has drastically dropped.
Despite the crisis of the biodiesel sector, mainly due to the origin,
production and supply of the raw oily material for such process, the
glycerol production is expected to rise and applications are sought
in a variety of elds, including the energy and chemical sectors.
Thus, glycerol is regarded as a building block for the production
of a variety of chemicals, including glycols and epychlorhydrin,
and applications of such polyalcohol will benet from the current
market situation [13].
The esterication of glycerol and the carboxylic acids from rosin
(mainly abietic acid) is a very usual reaction in industry (for up to
8090 years) [1,3]. Even in the latest years, hundreds of patents
regarding the production of rosin triglyceride and its applications
have been registered.
Rosin triglyceride is a standardized product, with a series of
ASTM-based standards that the product must full before its commercialization. Those ASTM standards refer to its colour (Gardner
and USRG colour scales), its viscosity behaviour with temperature (softening-point, ASTM-E 28-92), its acid composition
(acid number, ASTM-D 465-92), the presence of inorganics or
ash (ASTM-D 1063-51), the saponicable compounds to control
decarboxilation during the esterication process (saponication
number, ASTM-D 464-92), among others. In contrast to the classical analytical methodology used in the industrial standards, the
qualitative and quantitative chemical analysis of rosin and its products also include the newest technologies nowadays: SPE and HPLC,
MALDI-TOF, and capillary electrophoresis [1416].
However, knowledge regarding the process kinetics, i.e. the production rate of the involved chemical species and its mathematical
modelling, is scarce. In fact, information in literature refers to the
evolution of acidity (acid number) with time, with some studies
on the applied kinetics of similar esterications [1720]. In these
works, kinetic information refers to other acids used (including
fatty acids) as well as some other polyalcohols. In any case, potential
kinetic models are selected. For esterication of simple acids and
alcohols, a potential second-order kinetic model with partial rstorder for alcohol and acid, respectively, was proposed. For fatty
acids, this model is considered in some cases, whereas, in other
cases, a rst-order kinetic model is also used to t experimental
data [20].
Recently, Salmi et al. [21] worked on data obtained by Flory [18]
for the reaction between adipic acid and ethyleneglycol. They proposed a potential kinetic model whose total order depends on the
conversion of the acid. In fact, it depends on the acid concentration,
as the authors claimed, just like the proposed reaction mechanism.
At high acid concentration, the carbocation is formed in a relatively
polar liquid, and is thus stabilized: the order with respect to the
acid is one. At lower concentration, this intermediate is not stable
as the reaction medium is non-polar and ionic pairs are formed,
instead of the carbocation. Thus, a second order with respect to the
acid has to be considered. Therefore, the total order changes from
two to three as the reaction proceeds. The model also includes an
empirical parameter, q, which is related to the number of hydroxyl

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

groups of the alcohol being used (in this work, both lauric alcohol
and ethyleneglycol were considered) and, moreover, to the polarity
of the alcohols. The higher the value of this parameter, the longer
it takes to change from second to third order in the kinetic model.
Smith and Eliot [22] proposed a kinetic model for the esterication of rosin acids with lauric alcohol at 230260 C and
pentaerythritol at 260300 C. The kinetic model considered a partial second order for rosin, accounting only for the evolution of such
compound in the rst reacting system. When pentaerythritol was
used, the order depended somehow on temperature, being second
order at low temperatures and third order at the highest values. The
activation energy values, considering second-order kinetic models, were 64 KJ mol1 and 88 KJ mol1 , for the lauric alcoholrosin
system and for the pentaerythritolrosin system, respectively.
In this work, the kinetics of the esterication of rosin acids
and glycerol are studied and used to simulate industrial fed-batch
operation runs. For such purpose, a rst step was to develop an
automatic and robust analytical method based on gel permeation
chromatography, more informative but able to be used instead
of the classic acid number titration method (ASTM-D 465-92). To
study the evolution of glycerol in such a complex system, H NMR
and ionic chromatography methods were developed and employed.
After studying the temporal evolution of the involved chemical
compounds, kinetic models (rst-order, second-order and a hyperbolic model) coupled with different mass balances for glycerol
(with or without stripping of glycerol) were proposed, being the
models tted to acid conversion data utilized to select the most
adequate one. Finally, considering the results on the evolution of
glycerol concentration with time in the liquid phase and in the
distillate, as well as the temporal evolution of temperature and
glycerol addition during the industrial runs, a model that takes into
account the reaction kinetics and the stripping of glycerol was used
to simulate industrial operations.
2. Experimental
2.1. Materials
Rosin and glycerol were of technical and pharmaceutical grade,
respectively, and were kindly supplied by LURESA. Deuterated
water and chloroform for glycerol analysis by H NMR were purchased to SDS. Tetrahydrofuran for the GPCHPLC analysis of rosin
acids and their glycerides was purchased to SDS, too.
2.2. Esterication runs and sample analysis
The kinetic runs were performed between 240 and 280 C,
changing the initial concentration of glycerol between 9 and 15%
based on rosin weight (glycerol concentration = 0.91.5 mol L1 ).
In a typical isothermal run, 100 g of rosin were charged into a
500 mL three-necked round-bottom ask with upper agitation by
a marine helix and a distillation head attached so as to totally
condensate the water from the reaction together with trementine
(which contains mainly pinenes), thus the distillate volume can
be accurately measured (the set-up is depicted in Fig. 1). When
the temperature of reaction was reached, a time zero sample was
withdrawn with a glass pipette and glycerol was added. During the
reaction time other samples were withdrawn and stored in aluminium capsules. These samples were analysed by titration with
KOH and using GPCHPLC on a Phenogel 300 7.5 column for low
molecular weight compounds (5 m particle diameter, 50 A pore
diameter) using tetrahydrofuran as eluent and a DAD-detector at
254 nm wavelength. The direct basic titration used is based on the
ASTM D-465 standard and uses a freshly made KOH 0.06 N solution in ethanol standardized with a 0.1 M HCl solution, to titrate

321

Fig. 1. Experimental set-up for rosin and glycerol esterication at lab and pilot plant
scales.

0.3 g of the sample solubilised in 10 mL toluene with a drop of phenolphthalein (10 g L1 in ethanol). The reaction samples, as well as
distillate samples, were also analysed by H NMR using a BRUKER
AVANCE DPX 300 Mhz-BACS60 spectrometer. Glycerol content in
distillates was also analysed by ionic-GPC HPLC with RI detection of
compounds using a H+ -Rezex column for acids as stationary phase
and acidied mille-Q water (pH 2.2) as eluent.
In a typical temperature gradient run at laboratory scale, the
entire procedure was similar to the one used in isothermal runs
with some changes. On one hand, the equipment or set-up used was
larger, using a cylindrical reactor able to contain up to 1 L reaction
medium with a thermostated vessel located above for the addition
of glycerol. On the other hand, a temperature gradient simulating
the ones used at industrial scale was programmed by means of
the same PID temperature controller (E5CN of OMRON electronics) used for the isothermal runs. When a temperature of 240 C
was reached, a time zero sample was withdrawn using a valve at
the bottom of the reactor and glycerol was added at 230250 C.
All the samples withdrawn during the course of the run were analysed by GPCHPLC and by H NMR. Temperature was also monitored
during the run. Industrial runs at LURESA were carried out in a
similar manner and samples were analysed by the mentioned titration procedure (an industrial standard test) and some of them by
GPCHPLC.
2.3. Kinetic modelling and simulation
Several kinetic models were tted to experimental data from the
isothermal runs: order 1, order 2 and the hyperbolic model. The rst
order model is a pseudo-rst order kinetic model that considers a
constant concentration of the glycerol in the reaction liquid. The
second order kinetic model is rst partial order with respect to
the alcohol and to the acid, as is typical in esterication reactions.
The hyperbolic model represents an intermediate situation where
the kinetics of the reacting system follows rst-order behaviour at
the beginning of the reaction and evolves towards a second-order
model at the end of each run. Coupled with the kinetic model, three
mass balances for glycerol have been considered: in the rst one,
glycerol concentration decreases in the liquid phase only due to the

322

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

esterication reaction; in the second mass balance, a second term


taking into account stripping of glycerol due to evaporation of water
(product of the reaction) is added. In the latter case, the amount
of stripped glycerol is proportional to the rate of reaction. Finally,
in the last mass balance, glycerol concentration decreases due to
chemical reactions and to a stripping phenomenon proportional
to both reaction rate and the concentration of glycerol at a given
reaction time.
The tting was performed using the MarquardtLevenberg
algorithm coupled with a Runge-Kutta method for the numerical integration of the kinetic equations. The selection of the most
appropriate model was based on usual physical (positive value of
the kinetic parameters and values for the activation energies within
adequate ranges) and statistical criteria (Standard error value for
each kinetic parameter, Fischers F, Akaikes modied information
criterion and SQR values for each kinetic model).
For the non-isothermal (transient) runs, simulations with the
most adequate kinetic model were performed considering the
temporal evolution of the liquid temperature. Depending on the
evolution of the temperature during the run, the temperature values were tted by a hyperbolic or a sigmoid (Boltzmann) function
to take into account this variable in each simulation. Simulations
were performed considering both a fast addition of glycerol (for the
run performed in the laboratory) and a linear addition with respect
to time (for industrial runs). Moreover, and considering the results
regarding glycerol in lab scale isothermal runs, simulations of transient runs at laboratory and industrial scale included not only the
kinetics of the chemical reaction but also the stripping of glycerol
due to the production of water during esterication.

3. Results and discussion


3.1. Temporal evolution of isothermal runs analysed by
GPCHPLC and H NMR
Runs were performed in a batch reactor whose temperature
was controlled by a PID device. When the system was stabilized at
the working temperature, being the rosin totally melted, glycerol
was added and samples were withdrawn from the liquid medium,
while turpentine remaining in the rosin and the water formed during esterication condensed and were collected in a graduated test
tube. Due to the low solubility of turpentine in water, the latter
was easily separated and analysed by H NMR and ionic HPLC, and
the samples taken from the reacting liquid were analysed by H
NMR. The distillate was analysed by H NMR using deuterated water
and the samples withdrawn from the rosin-glycerol mixture were
dissolved in deuterated chloroform.
Analyses on the distillates by H NMR showed that a very considerable amount of glycerol was present on the water. When
subjecting these samples to ionic HPLC analysis, a distinct glycerol
peak was observed and glycerol was quantied by using the adequate calibration equation and knowing the amount of collected
water. In the case of H NMR spectra, a signal at = 5.1 ppm can be
ascribed undoubtedly to hydrogen atoms in water and alcohols,
while a very rough signal is observed at = 3.43.5 ppm. The latter
signal can be in rst instance ascribed to hydrogen atoms linked to
the methylene and methylene carbons in the molecule of glycerol.
The Signal is so rough, apparently, due to coupling of the different
hydrogen atoms in adjacent carbon atoms. Quantication of glycerol concentration was performed in this case from the values of
the areas of both signals, correcting the area of signal at = 5.1 ppm
by subtracting the amount of this area due to hydrogen atoms in
glycerol hydroxyl moieties. All calculations were performed on the
base of one hydrogen atom (thus, on a molar base for water and for
glycerol) and, thus, the apparent concentration of glycerol in water

was calculated. When comparing these results from H NMR analysis to the results from ionic HPLC, where a distinct peak clearly
due to glycerol is obtained, the concentration of glycerol seemed
to be 5060% higher in the former case. These results are probably due to the fact that turpentine is mainly composed by and -pinene, but it also contains alcohols, aldehydes and other
polar compounds. The mentioned compounds could be observed
in the chromatograms, where six peaks at longer times than that
of glycerol were present, though their signals were not signicant.
Hydrogen atoms of these compounds could give signal in H NMR at
= 3.43.5 ppm, so that concentration of glycerol based on H NMR
analysis is higher than that calculated from HPLC analysis of the
distillates. As a consequence, the percentage of glycerol in the distillates stripped during the runs was calculated in each of them
using the ionic HPLC analysis. Such results are shown in the last
column in Table 1.
The H NMR analysis of the samples from the liquid in the reactor was used to quantify the conversion of glycerol in the reacting
mixture by using the following equation:
XG = 1

AG ( = 3.5ppm)
AG ( = 3.5ppm) + AE ( = 3.7ppm)

(1)

This equation considers the signals due to hydrogen atoms in


all carbons of the glycerol molecule (signal at = 3.5 ppm) and to
the methylene carbons in the ester molecules. Esterication of the
hydroxyl groups in such positions leads to the decrease of the signal at = 3.5 ppm with the concomitant increase of the signal at
3.7 ppm. However, signal at = 3.5 ppm cannot be ascribed without doubt to glycerol and, thus, nal values for the conversion of
glycerol are only of a semi-quantitative nature. Moreover, quantication of glycerol signal at high conversion values is prone to be
erroneous due to the low ratio signal/noise in H NMR spectra.
NMR techniques have also been employed to observe whether
acrolein was present in distillates, since the high temperatures
used in the runs could be adequate for its production, even if no
strong acid catalyst is used. 13 C and H NMR spectra of distillates
in D2 O and deuterated DMSO show evidence of the inexistence of
this compound, as there are no signals at high displacement values
( = 67 ppm and = 9.6 ppm in the H NMR DMSO spectrum, and
= 138 ppm and = 196 ppm in the 13 C NMR DMSO spectrum), as
expected if acrolein is formed, stripped and condensed.
The other analytical tool used is based on the separation of the
esters in a gel permeation chromatographic column of a very nar Such procedure leads to separation of the
row pore diameter (50 A).
esters and the rosin acids (considering the latter as a whole: abietic,
pimaric and their isomers). In Fig. 2, a comparison of traditional
standards of polystyrene used in these kinds of chromatography and rosin and its esterication products is made and a good
agreement in elution time (and volume) is observed between the
standards and the analytes, considering their molecular weights.
UVvis-spectrophotometric scanning of several rosin acid standards resulted in the adscription of the signal at 254 nm to abietic
acid and its isomers; therefore the absorbance at this wavelength
is due to the terpenic moiety of the different rosin acids, neither
to the carboxylic acid group nor to the ester bond [1]. Considering this fact Fig. 3 shows the percentage of rosin acids in its acid
form and being part of the several esters in the mixture as these
compounds appear and disappear from the reaction mixture with
time. The run herein shown was conducted at 240 C and the result
shows that rosin evolves towards its triglyceride rapidly, with little
increase of the diglyceride and an even lower amount of monoglyceride in the mixture. In fact, at higher temperatures, the amounts
of monoglyceride and diglyceride were even lower. Thus, in industrial operation (at 275 C), these esters are in minor concentration.
It should be taken into account that the conversion of rosin has
a similar trend to the conversion of glycerol, on one hand, and

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

323

Table 1
Isothermal runs at lab scale: conditions, rosin and glycerol conversion at 240 min reaction time and stripped glycerol (in percentage of the initial amount of the polyalcohol).
Run

CGo (M)

T0 ( C)

XC 240

XG 240

Stripped glycerol (% w/w)

1
2
3
4
5
6
7
8
9
10
11
12

1.5
1.3
1.1
1.0
0.9
0.7
1.0
1.0
1.0
1.0
1.0
1.0

0.50
0.43
0.37
0.33
0.30
0.23
0.33
0.33
0.33
0.33
0.33
0.33

270
270
270
270
270
270
240
260
265
270
275
280

0.91
0.84
0.82
0.83
0.77
0.62
0.64
0.80
0.81
0.82
0.84
0.91

0.73
0.80
0.87
0.88
0.88
0.93
0.68
0.83
0.85
0.88
0.90
0.99

12.4
11.4
10.7
9.6
7.7
6.5
8.0
6.1
9.1
11.3
9.5
10.7

1200

+10%
1

polystyrene standards
rosin and its esters

-10%

0.9
0.8

800

0.7

600
Xc by titration

Molecular weight (D)

1000

400
200
0

0.6
0.5
0.4
0.3

10

11

12

13
0.2

Elution time (min)


Fig. 2. HPLCGPC standard curve of rosin and rosin esters in the Phenogel 50 A
column: comparison to GPC polystyrene standards.

0.1
0
0

100

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Xc by GPC-HPLC

monoglyceride
Fig. 4. Comparison of the rosin conversion data obtained by the traditional acid
number method (ASTM D 465-92) and those obtained by the HPLCGPC method.

diglyceride
triglyceride

80
% weight

0.1

rosin

60

40

20

0
0

50

100

150

200

250

300

350

400

rosin triglyceride is formed directly from 3 mol of rosin and 1 mol


of glycerol.
When comparing the areas of the peaks in the chromatograms
to the result of the titration using the acid number standard, a
relatively good agreement was observed (data not shown). When
all the areas of the peaks corresponding to the compounds of interest where summed and each area was divided by the total sum,
relative areas (%A) were calculated and they were in good agreement with the acid measured by titration (as shown in Fig. 4). Thus,
GPCHPLC is a robust and automatic method eligible to utilization
instead of the more cumbersome titration analysis, and the data
thence obtained can be used for further statistical analysis. Thus,
the conversion of rosin was calculated as:

time (min)
Fig. 3. Temporal evolution of weight percentage of rosin acids and esters at M = 0.30
and T = 240 C.

that concentrations of intermediates are low and almost constant


(especially in the case of monoglyceride). The application of the
quasi-steady state hypothesis to monoglycerides and diglycerides
shows that all reactions take place at the rate of the rst reaction,
being the latter the controlling step. In this situation, the kinetic
scheme of the reacting system can be simplied from a scheme with
three-reactions in series to a one reaction scheme where 1 mol of

XC = 1

%Ac (t)
%Ac (t = 0)

(2)

Results from the runs performed under isothermal conditions,


at several glycerol concentrations and temperatures, are shown in
Table 1. Herein, the conversion of rosin and glycerol at 240 min can
be seen for all the runs, being the conversion of glycerol obtained
from H NMR analysis of samples from the reacting liquid and the
percentage of stripped glycerol to the glycerol added at time zero
calculated from ionic HPLC measurements. Though it is not shown,
conversion values were calculated for all the withdrawn samples.
Regarding the conversion of rosin (only due to esterication reac-

324

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

tions) and that of glycerol, glycerol concentration in the reacting


liquid decreases at a higher rate than expected because of the esterication process, by means of which the evaporating water being
produced strips off some of this polyalcohol from the reacting liquid. This phenomenon seems to be in according with the initial
concentration of glycerol: the higher it is, the higher the percentage
of stripped glycerol. This fact, together with the necessity of reaching acid numbers below 7 in the nal product, explains why the
common industrial practice is to add more glycerol than that stoichiometrically needed (a 3040% in excess to the 1:3 glycerol:rosin
molar ratio).
3.2. Kinetic modelling of the esterication reaction
As the evolution of the reacting mixture can be well represented
by the evolution of the acid, data from the GPCHPLC analysis concerning the mixture of rosin acids (peak at 9.8 min elution time)
were corrected relating its area to the sum of areas of all the
involved compounds. These data were thereafter used to t three
kinetic models:
Model A: A rst-order model, which is quite usual in the esterication of glycerol and fatty acids.
r=

1 dCC
= kCC
3 dt

(3)

Model B: A second-order model with partial rst-order both for


the alcohol and the acid, commonly used to model the esterication of monoacids and monoalcohols.
r=

1 dCC
= k CG CC
3 dt

(4)

Model C: A hyperbolic model whose numerator corresponds to


the second-order model, and whose denominator includes the
concentration of glycerol. It may be simplied to one of the aforementioned models at an adequate glycerol concentration.
r=

k1 CG CC
1 dCC
=
1 + K2 CG
3 dt

(5)

The second order potential model and the hyperbolic model were
coupled with mass balances of glycerol where its concentration
changes with time considering three possibilities:
Mass balance A: Glycerol concentration decreases only due to the
chemical reaction in which it is involved. Thus, glycerol evolution
rate is a third of rosin evolution rate.
1 dCC
dCG
=
3 dt
dt

(6)

Mass balance B: The decrease in glycerol concentration is due to


the reaction rate and stripping of the polyalcohol driven by the
water formed during the reaction. Thus, a second term directly
proportional to rosin disappearance rate is added.
dCG
1 dCC
=
(1 + kst )
3 dt
dt

(7)

Mass balance C: The third possibility considers that the stripping


of the polyalcohol not only is proportional to the rosin disappearance rate, but also to the initial concentration of glycerol.
1 dCC
dCG
(1 + kst CG )
=
3 dt
dt

(8)

Combinations of the kinetic models and glycerol mass balances


yield the models that have been tted to the experimental data.
A rst set of runs (16) was performed at 270 C changing the
ratio of the polyalcohol to the major component, rosin. In a rst

instance, all runs were tted to reach optimised values of the


kinetic parameters for each model at 270 C. As shown in Table 2,
all parameters were sound from a statistical point of view, having
all standard errors smaller than the parameter value. The sum of
residual squares (SQR) decreased steeply from the rst order model
without stripping of glycerol to the hyperbolic model with stripping of glycerol proportional to rosin global disappearance rate and
glycerol initial concentration. However, the number of parameters
changes from one to three, so SQR decrease is also inuenced by this
fact. Thus, two other criteria were used to select the best model:
Fischers F95 value (based on a null hypothesis on the suitability of
the modelmass balance to t data) and Akaikes information criterion corrected for a ratio number of data to number of parameters
<40 (AICc). The Fischers F value is a null hypothesis analysis at 95%
condence which has been applied in the discrimination of kinetic
models (and in regression techniques in general) for a long time. In
this work, F95 value is calculated by Eq. (9), hence a high number
of parameters implies a great reduction in its value:
F95 =

SSQWPV/K
SQR/ (N K)

(9)

The Akaikes information criterion also places a penalty on the


use of excessive parameters in any model used to t data. Information criteria were created in the rst seventies [23] and have
continuously been developed, probed into and applied in a variety of elds in the last twenty years [2426]. The original Akaikes
information criterion (AIC) was devised for very large sets of data
and can be calculated from:
AIC = NLn

 SQR 
N

+ 2K

(10)

However, when the ratio of number of data to number of parameter is lower than 40, it is better to use a modied version of this
information criteria, the AICc (Saha et al., 2008), according to:
AICc = AIC +

2K (K + 1)
NK 1

(11)

The higher the F and the lower the AICc parameter values, the
more prone the model is to be correct. In Table 2, the F95 value suggests that the best model is the hyperbolic model without stripping
of glycerol (model 5), while the corrected Akaikes information criterion supports hyperbolic models with stripping of glycerol, more
in accordance with the existence of glycerol in the distillates (models 6 and 7).
For further discrimination among the proposed kinetic models
and glycerol mass balances, a second set of isothermal runs (712)
was executed, with temperatures ranging from 240 C (when glycerol is usually added in an industrial run) to 280 C (the upper
temperature in industrial practice). Afterwards, data from all the
runs were used in a multivariable tting to select the most appropriate kinetic model with the same criteria used in the previous
set. As it can be seen in Table 3, all the tested models are reasonable considering the standard errors of their parameters and the
F value (always higher than that needed at 95% condence for the
given number of data and parameters), while the AICc value aids in
the selection of the most adequate one. However, the model which
again best ts the experimental data is the hyperbolic model, whose
constant (K2 ) in the denominator is not temperature dependent. In
this model, glycerol is stripped from the reaction liquid at a rate
proportional to rosin disappearance rate and to the concentration
of glycerol (model 7). When tting this model to all data, both the
highest value for F and the lowest value for the AICc parameters
were obtained. Fig. 5ac depicts the tting of the selected model
(lines) to the experimental data (points), proving to be an adequate
tting for all the tested conditions.
The hyperbolic model considers a situation in all runs where
at short times the reaction evolves following rst-order kinetics,

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

325

Table 2
Comparison among the proposed kinetic models: kinetic and statistical parameters at a given temperature (270 C).
Runs 16 (M = 0.250.5; T = 270 C)
Model reference-equations

Parameters

Value error

SQR

AICc

F95

Model 1 (Eqs. (3) and (6))


Model 2 (Eqs. (4) and (6))
Model 3 (Eqs. (4) and (7))

k
k 
k 
kst
k 
kst
k1
K2
k1
K2
kst
k1
K2
kst

3.33 103 1.34 104


4.98 103 1.36 104
4.04 103 1.67 104
5.98 102 3.35 104
4.55 103 2.17 104
7.84 102 5.53 104
7.61 103 1.18 103
1.01 2.38 101
5.99 103 5.23 104
4.28 101 1.01 101
1.13 101 1.63 102
5.58 103 4.09 104
4.28 101 1.01 101
2.37 101 3.51 102

0.813
0.315
0.235

4.75
5.70
5.89

70
160
105

0.239

5.87

104

0.155

6.31

163

0.119

6.42

137

0.119

6.42

136

F95

Model 4 (Eqs. (4) and (8))


Model 5 (Eqs. (5) and (6))
Model 6 (Eqs. (5) and (7))

Model 7 (Eqs. (5) and (8))

Note: k (min1 ); k  , k1 (L mol1 min1 ); K2 (L mol1 ); kst (dimensionless).


Table 3
Comparison among the proposed kinetic models: kinetic and statistical parameters from multi-temperature tting.
Runs 16 (M = 0.250.5; T = 270 C)
Model reference-equations
Model 1 (Eqs. (3) and (6))
Model 2 (Eqs. (4) and (6))
Model 3 (Eqs. (4) and (7))

Model 4 (Eqs. (4) and (8))

Model 5A (Eqs. (5) and (6))

Model 5B (Eqs. (5) and (6))

Model 6 (Eqs. (5) and (7))

Model 7 (Eqs. (5) and (8))

Runs 712 (M = 0.33; T = 240280 C)


Parameters


ln k 0
Ea
ln k  0
Ea (J mol1 )
ln k  0
Ea
kst
ln k  0
Ea
kst
ln k10
Ea
K2
ln k10
Ea1
ln K20
Ea2
ln k10
Ea
K2
kst
ln k10
Ea
K2
kst

Value error

SQR

AICc

6.98 1.41
6579 925
7.09 1.08
6678 578
7.20 1.08
6683 592
1.84 102 2.25 102
8.17 1.11
6733 599
6.21 103 3.81 102
7.68 0.83
6723 448
7.73 101 1.55 101
5.17 4.62
5241 2620
7.40 11.26
3878 6702
8.14 0.69
7209 374
3.87 101 6.47 102
1.32 101 1.02 102
8.22 0.67
7268 364
3.39 101 5.65 102
2.81 101 2.11 102

1.47

4.59

58

0.33

6.08

225

0.32

6.02

144

0.32

6.02

143

0.25

6.27

209

0.24

6.19

127

0.16

6.60

238

0.15

6.66

253

Note: Ea (J mol1 ); K2 (L mol1 ); T (K); kst (dimensionless).

and at longer times second-order kinetics is able to t well data. In


fact, the rst order model constant shows an inverse trend when
compared to the second order one, while no trend with glycerol
concentration is observed in the hyperbolic model numerator constant. In this kind of models, the denominator constant usually
accounts for a physical phenomenon: adsorption in LHHW models,
intermediate complex formation in MichaelisMenten kinetics, to
mention a couple. In this case, where the system is not catalytic
by nature, the constant K2 in the denominator could be reecting
mass transfer or solubilization phenomena, being this an empirical
parameter (thus, it is not deduced from any reaction mechanism).
In fact, rosin acids and glycerol, as well as other alcohols usually employed to produce esters of such acids may be not totally
miscible with each other due to their difference in polarity. The
hyperbolic model suggests that, at the initial reaction period, rst
order with respect to rosin is enough to t data, while at longer
reaction times a second order model is needed. The second order
model (rst order with respect to each reagent) is the classical
kinetic model in esterication when both reagents are in the same

phase. The rst order model suggest that glycerol concentration is


not changing at the beginning of the reaction, so the liquid phase
where the reaction takes place seems to be saturated with glycerol.
In this situation, a high glycerol concentration at the beginning of
each run implies that the second term in the denominator prevails
over the rst one. This means that not all of the added glycerol is
solubilized into the rosin phase, where the reaction proceeds. Actually, as shown by the evolution of monoglycerides and diglycerides,
once the hydroxyl moieties are solubilized into the acid phase, reactions occur at a very fast rate, yielding triglycerides, especially in
what concerns monoglycerides. Thus, the hyperbolic model leads
to a better t because it can reect empirically this transition from
a pseudo-rst order model to the classical second order model.
The parameter in the denominator, K2 , could reect the solubility of the glycerol into the rosin phase. Its variation with
temperature supports this, as ttings of hyperbolic kinetic models
considering or not change with temperature to experimental data
strongly suggest that this parameter is not a function of temperature (higher value of F, lower value of AICc). While kinetic constants

326

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

1.0

for chemical reactions, adsorption, and mass transfer phenomena


are usually inuenced by temperature, solubility changes with temperature strongly depend on the nature of both the solute and the
solvent; therefore, solubility could be independent of temperature,
at least under certain conditions.
The stripping of glycerol is observed in the last column of Table 1,
and considered in two of the coupled hyperbolic model-glycerol
mass balance equations. Considering the results in Table 1, it is
obvious that glycerol stripping is a function of the initial concentration of polyalcohol being added. From results in Tables 2 and 3,
the value of the parameter kst in the second glycerol mass balance
compared to unity is similar to the fraction of glycerol stripped calculated by HPLC analysis of distillates. In fact, the product of the
parameter kst calculated for the third glycerol mass balance and
the initial concentration of glycerol changes with this latter variable similarly to the experimental values. This is a physical probe
that the selected kinetic model and to the selected mass balance
for glycerol are adequate for the system studied.

0.9
0.8
0.7

Xc

0.6
0.5
0.4

run 1-M=0.50

0.3

run 2-M=0.43

0.2

run 3-M=0.37

0.1

model fitting

0.0
0

200

400

600

800
3.3. Simulation of industrial runs with the selected kinetic model

time (min)

After isothermal runs in the set-up shown in Fig. 1 were carried


out and analysed, some runs (shown in Table 4) were performed
using the capacity of the PID reactor to control heating following
an almost linear variation in the liquid temperature: run 13 is an
example of such runs. This kind of operation is similar to that followed at industrial scale, to minimize the energy and raw material
costs. In fact, runs 1416 are examples of industrial runs and follow this non-stationary to stationary operation, with addition of
glycerol at 240 C. Thus, to simulate this operation, two steps are
needed: in the rst phase, a semi-batch operation has to be considered; a batch operation is considered from the time when the nal
operating temperature is reached. In the rst step, as for the runs at
lab scale, the addition of glycerol can be considered instantaneous,
while a linear addition with respect to time during a given period
has to be considered for the industrial runs. Moreover, in this step,
the simulation has to take into account the temporal evolution of
temperature, as shown in Figs. 5a and 6a. This evolution resulted
in following sigmoid or hyperbolic changes with time, as given by:

1.0

0.9
0.8
0.7

Xc

0.6
0.5
0.4
0.3

run 4-M=0.33

0.2

run 5-M=0.30
run 6-M=0.23

0.1

model fitting

0.0
0

200

400

600

800

time (min)

Sigmoid (Boltzman) T ( C) =
Modied hyperbole :

0.9

0.7

0 = RC +

0.6

Xc

T ( C) =

1
(t(h) + 2 ) + 3

(12)
(13)

Further complication of the simulation has to consider stripping of glycerol as shown in the previous section. Thus, in the rst
step of the operation, rosin concentration evolves according to the
following mass balance:

0.8

0.4
run 7-T= 280 C

0.3

run 10-T= 265 C


0.2
run 12-T= 240 C
0.1

FG
dCG
= RG +
V
dt

model fitting

0
100

200

300

400

500

600

700

dCC
dCC
k1 CG CC
= 3r +
r=
(1 + kst CG )
1 + K2 CG
dt
dt

(14)

where K2 has no dependence on temperature and k1 follows the


Arrhenius equation with the parameters shown in Table 3. For the
rst step of the run, the temperature follows an empirical function
obtained for each particular run and is included in the simulation
together with Eq. (5). Meanwhile, glycerol concentration changes,
during glycerol addition, according to:

0.5

2 + (1 2 )
(1 + exp((t(h) t0 )/dx)

800

time (min)
Fig. 5. Fitting of the hyperbolic kinetic model (lines) to experimental esterication
data (points). At T = 270 C and M = 0.37, 0.43, 0.50. At T = 270 C and M = 0.23, 0.30,
0.33. At T = 240, 265, 280 C and M = 0.33.

(15)

Being the addition term (FG ) equal to zero after the addition step.
In every step, with or without addition of glycerol, the conversion
of rosin can be estimated by:
XC =

nC0 nC
nC0

(16)

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

327

Table 4
Non isothermal runs (transient temperature) at pilot plant and industrial scale: operational conditions and values of the stripping coefcient.
Run

Scale

nC0 (kmol)

FG (mol min1 )

 (h)

T0 ( C)

kst

13
14
15
16

Pilot plant
Industry
Industry
Industry

0.001
26.2
26.6
27.6

63
94
98

3
2
2

180
255.5
240
225.7

4.91 102 1.03 102


4.13 101 5.62 102
4.01 101 6.15 102
3.32 101 28 102

290

280

270

260

230

T (C)

T (C)

250

210

240

190

T profile data run 14

T profile
170

T profile data run 15

220

Boltzmann function fit

T profile data run 16

150

T profile fitting

50

100

150

200

250

300

200

time (min)

200

400

600

800

1000

time (min)

1.0
0.9

1.00

0.90
0.80
0.70

0.5
0.4
0.3
0.2

0.60
Xc

Xc

0.8
0.7
0.6

0.1
0.0
0

50

100

150

0.50
Data run 14

data run 13

0.40

simulation run 13

0.30

Data run 16

0.20

Simulation run 14

200

250

300

time (min)
Fig. 6. Transient operation at laboratory pilot plant scale (run 13): (a) tting to
Boltzmann function of the temporal temperature prole; (b) simulation (line) of
experimental data (points) with the selected hyperbolic model and glycerol mass
balance.

To simplify the simulation, the volume of the reacting liquid has


been considered to be that of the added rosin. This is so because it
seems that reactions proceed in the rosin phase (at the beginning
of each run, it seems that not all the added glycerol is solubilized
into the rosin phase). Other reasons are the relatively low volume
of glycerol compared to that of rosin and the reduction of the total
volume due to water and turpentine evaporation, linked to stripping of some of the added glycerol. Thus, a maximum error of 5% is
introduced in the simulation due to this simplication.
In Fig. 6, the tting of temperature ( C) with time is shown
along with a simulation of transient temperature for run 13, at laboratory scale, where the selected couple kinetic modelglycerol
mass balance is taken into account. In this case, as in the runs
performed at industrial scale, the term accounting for the stripping of glycerol is calculated as the reactor is different from the

Data run 15

Simulation run 15

0.10

Simulation run 16

0.00
0

200

400

600

800

time (min)
Fig. 7. Transient operation at industrial scale: (a) tting of functions to the temporal temperature proles in runs 1416; (b) simulation (line) of experimental data
(points) with the selected hyperbolic model and glycerol mass balance.

former, as reux of condensates returning into the reaction liquid could depend on the reactor set up used in each occasion. For
the runs performed at industrial scale (in the plant of LURESA in
Coca, Segovia) identical simulations have been performed, as can
be seen in Fig. 7. In all cases, considering a stripping of glycerol
proportional to the glycerol concentration in the reacting liquid
and to the rate of consumption of rosin (or production of water)
led to an adequate simulation of the transient process. For all the
cases, results are shown in Table 4, showing that little stripping
of glycerol is observed in the laboratory bench pilot plant, while
this phenomenon is much more important in the industrial reactor. This could be due to the fact that in the pilot plant reactor, reux
of condensates is partly allowed, while no reux is observed in the

328

M. Ladero et al. / Chemical Engineering Journal 169 (2011) 319328

industrial reactor (only a 20% free volume over the reacting liquid
and no condenser attached to the head of the reactor).
4. Conclusions
The kinetic study of the esterication of rosin acids and glycerol was performed at laboratory and industrial scale. For obtaining
wider information of the temporal evolution of chemical species
in the reaction mixture, NMR, ionic HPLC and GPCHPLC analytical methods have been successfully employed. The latter, when
using relative areas of the peaks of rosin acids and esters, has
proved that the amount of mono- and di-ester species in the liquid
was always low enough to consider that glycerol evolves directly
to triglycerides. Moreover, when compared to the standard acid
number titration method, GPCHPLC has proved to be robust
enough, while being more automatic and, thus, less prone to experimental error.
A set of runs at several glycerol to rosin molar ratio values and
temperatures have allowed to select a hyperbolic kinetic model
coupled with a glycerol mass balance in which stripping of glycerol occurs after a discrimination between several kinetic models
and glycerol mass balances. In the selected glycerol mass balance,
stripping of the polyalcohol was proportional to reaction rate and
to the concentration of the polyalcohol at a given time. Physical and
statistical criteria (standard error of parameters, Fischers F95 and
Akaikes AICc) were used for such discrimination.
Afterwards, transient temperature runs have been performed
at lab and industrial scale and tting of the selected model (leaving kst as tting parameter) led to their adequate simulations. At
lab scale, glycerol stripping was reduced, while at industrial scale,
stripping of glycerol was of importance. This could be due to reux
of condensates inside each reactor set-up.
Acknowledgments
Financial support from the Spanish Ministry of Education
and from LURESA (through project PETRI 95-0821.OP) is gratefully acknowledged. Moreover, the authors want to extent this
recognition to the Spanish Ministry of Environment (project MMA0392063-11.2) for further nancial support.
References
[1] E.J. Soltes, D.F. Zinkel, Chemistry of rosin, in: D.F. Zinkel, J. Russell (Eds.), Naval
Stores: Production, Chemistry, Utilization, Pulp Chemicals Association, New
York, 1989, pp. 261320.
[2] S. Zhaobang, Production and Standards for Chemican Non-Wood Forest Products in China, CIFOR report no. 6, 1995, ISSN 0854-9818.

[3] D.F. Zinkel, J. Russell (Eds.), Naval Stores: Production, Chemistry, Utilization.
Section 5: Rosin Utilization, Pulp Chemicals Association, New York, 1989.
[4] V.T. Dhanorkar, R.S. Gawande, B.B. Gogte, A.K. Dorle, Development and characterization of rosin-based polymer and its application as a cream base, J.
Cosmetic Sci. 53 (2002) 199208.
[5] C. Li, W. Li, H. Wang, D. Zhang, Z. Li, Effects of rosin-type nucleating agent and
low density polyethylene on the crystallization process of polypropylene, J.
Appl. Polym. Sci. 88 (2003) 28042809.
[6] S.V. Fulzele, P.M. Satturwar, A.K. Dorle, Polymerized rosin: novel lm forming
polymer for drug delivery, Int. J. Pharma 249 (2002) 175184.
[7] N. Yuanmei, Y. Xingdong, L. Fuhou, Sonochemical synthesis of maleated rosin,
Chin. J. Chem. Eng. 16 (2008) 365368.
[8] S. Liu, C. Xie, S. Yu, F. Liu, Dimerization of rosin using BrnstedLewis acidic
ionic liquid as catalyst, Catal. Commun. 9 (2008) 20302034.
[9] L. Wang, X. Chen, J. Liang, Y. Chen, X. Pu, Z. Tong, Kinetics of the catalytic isomerization and disproportionation of rosin over carbon-supported palladium,
Chem. Eng. J. 152 (2009) 242250.
[10] Y. Caili, Z. Faai, Preparation and characterization of rosin glycerin ester and its
bromide, Front. Chem. China 2 (2006) 158160.
[11] F. Wang, T. Kitaoka, H. Tanaka, Supramolecular structure and sizing performance of rosin based emulsion size microparticles, Colloids Surf. A:
Physicochem. Eng. Aspects 221 (2003) 1928.
[12] F. Aran-Ais, A.M. Torro-Palau, A.C. Orgiles-Barcelo, J.M. Martin Martinez, Synthesis and characterization of new thermoplastic polyurethane adhesives
containing rosin resin as an internal tackier, J. Adhes. Sci. Technol. 14 (2000)
15571573.
[13] M. Pagliaro, M. Rossi, The future of glycerol: new uses of a versatile raw material,
in: RSC Green Chemistry Book Series, RSC Publishing, Cambridge, 2008.
[14] U. Nilsson, N. Berglund, F. Lindahl, S. Axelsson, T. Redeby, P. Lassen, A.-T. Karlberg, SPE and HPLC/UV of resin acids in colophonium-containing products, J.
Sep. Sci. 31 (2008) 27842790.
[15] A. Findeisen, V. Kolivoska, I. Kaml, W. Baatz, E. Kenndler, Analysis of diterpenoic
compounds in natural resins applied as binders in museum objects by capillary
electrophoresis, J. Chromatogr. A 1157 (2007) 454461.
[16] Y. Kumooka, Analysis of rosin and modied rosin esters in adhesives by matrixassisted laser desorption/ionization time-of-ight mass spectrometry (MALDITOF-MS), Forensic Sci. Int. 176 (2008) 111120.
[17] R.P. Carter, Rosin ester development, Ind. Eng. Chem. 37 (1945) 448465.
[18] P. Flory, Kinetics of condensation polymerization: the reaction of ethylene glycol with succinic acid, J. Am. Chem. Soc. 59 (1937) 466470.
[19] A.T. Williamson, C.N. Hinselwood, The kinetics of esterication. The reaction
between acetic acid and methyl alcohol catalysed by hydrions, J. Chem. Soc. 23
(1934) 587590.
[20] M. Berrios, J. Siles, M.A. Martin, A. Martin, A kinetic study of the esterication
of free fatty acids (FFA) in sunower oil, Fuel 86 (2007) 23832388.
[21] T. Salmi, E. Paatero, P. Nyholma, Kinetic model for the increase of reaction order
during polyesterication, Chem. Eng. Process. 43 (2004) 14871493.
[22] T.L. Smith, J.H. Elliot, The kinetics of esterication of resin acids, J. Am. Oil Chem.
Soc. 35 (1958) 692699.
[23] H. Akaike, Canonical correlation analysis of time series and the use of an information criterion, in: R.K. Mehra, D.G. Lainotis (Eds.), System Identication:
Advances and Case Studies. Academic Press, New York, NY. 1976, pp 2796.
[24] B. Saha, P. Chowdhry, A.K. Ghoshal, Al-MCM-41 catalyzed decomposition of
polypropylene and hybrid genetic algorithm for kinetics analysis, Appl. Catal.
B: Environ. 83 (2008) 265276.
[25] B. Saha, P.K. Reddy, A.K. Ghoshal, Hybrid genetic algorithm to nd the best
model and the globally optimized overall kinetics parameters for thermal
decomposition of plastics, Chem. Eng. J. 138 (2008) 2029.
[26] J.J. Knol, J.P.H. Linssen, M.A.J.S. van Boekel, Unravelling the kinetics of the formation of acrylamide in the Maillard reaction of fructose and asparagine by
multiresponse modelling, Food Chem. 120 (2010) 10471057.

Anda mungkin juga menyukai