Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Bioresource Technology 102 (2011) 3504–3511

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Disproportionation of rosin on an industrial Pd/C catalyst: Reaction pathway


and kinetic model discrimination
Juan Carlos Souto, Pedro Yustos, Miguel Ladero ⇑, Felix Garcia-Ochoa
Dpt. Ingeniería Química, Fac. CC. Químicas, Universidad Complutense, 28040 Madrid, Spain

a r t i c l e i n f o a b s t r a c t

Article history: In this work, a phenomenological study of the isomerisation and disproportionation of rosin acids using
Received 18 August 2010 an industrial 5% Pd on charcoal catalyst from 200 to 240 °C is carried out. Medium composition is deter-
Received in revised form 1 November 2010 mined by elemental microanalysis, GC–MS and GC-FID. Dehydrogenated and hydrogenated acid species
Accepted 3 November 2010
molar amounts in the final product show that dehydrogenation is the main reaction. Moreover, both
Available online 12 November 2010
hydrogen and non-hydrogen concentration considering kinetic models are fitted to experimental data
using a multivariable non-linear technique. Statistical discrimination among the proposed kinetic models
Keywords:
lead to the conclusion hydrogen considering models fit much better to experimental results. The final
Rosin acids
Disproportionation
kinetic model involves first-order isomerisation reactions of neoabietic and palustric acids to abietic acid,
Dehydrogenation first-order dehydrogenation and hydrogenation of this latter acid, and hydrogenation of pimaric acids.
Kinetics Hydrogenation reactions are partial first-order regarding the acid and hydrogen.
Pd/C catalyst Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction 200 to 240 °C. It is applied in several industries: adhesives (hot-


melt and pressure-sensitive adhesives), solder flux, printing inks,
Rosin is obtained from living trees (gum rosin), from pine paper neutral-size – after saponification – and more (Soltes and
stumps (wood rosin), and from the alkaline extraction of wood Zinkel, 1989). It has some important applications in the polymeric
during the Kraft pulping process (tall-oil rosin). Rosin is mainly a industry where abietic acid presence is not adequate, because it
mixture of C20 monocarboxylic diterpenic resin acids of the acts as an inhibitor (amounts as low as 0.5% in abietic acid are
abietic-type acids (mainly abietic, palustric and neoabietic acids) looked up) and very high percentages in dehydroabietic acid are
and of the pymaric acid type (pimaric and isopimaric acids, usu- desired (65% or more): as an emulsifier in the production of sty-
ally). It also has a certain quantity of fatty acids (in tall oil) and rene–butadiene rubber and ABS resin and chloroprene rubber
other neutral components (5–15%). The main difference between (Mayer et al., 1995, 1996). New fields where applications are being
rosin acids is a conjugated double bond system present in abietic developed include the synthesis of non-ionic surfactants (Hu et al.,
type resin acids, but not in the pimaric family. Carboxylic and 2006), applications in the coating industry in antifouling paints
olefinic functionalities of resin acids are advantageously modified and in the synthesis of new polymers (Duan et al., 2009).
by hydrogenation, disproportionation, formation of adducts with Dehydroabietic acid, as other abietates, is a diterpene acid
dicarboxylic acids, dimerisation, polymerisation, esterification important in the defense system of conifers, against potential her-
and saponification (Soltes and Zinkel, 1989). Some applications re- bivores and pathogens (González et al., 2010). Dehydroabietic acid
quire a combination of reactions to obtain the desired product (DHA) and its derivatives show a broad spectrum of biological ac-
quality. For example, esterification with polyols, disproportion- tion: antiulcer, antimicrobial, anxiolytic, antiviral, antitumor, and
ation and isomerisation reactions occur together, leading to esters cytotoxic activities (González et al., 2010; Tanaka et al., 2008;
of rosin acids stable towards oxidation and, thus, of a very clear Fonseca et al., 2004). Recently, DHA structure has been proved to
colour – one of the desired properties of rosin and rosin products. be a new scaffold for BK (large-conductance calcium-activated
Disproportionated rosin has good oxidation resistance, low K+) channel openers, causing a dramatic increase of the channel-
brittleness, high thermal stability and light color, and maintains opening activity. In fact, as it has shown one of the greatest activ-
a high softening point, even higher than the original rosin due to ities ever reported, it seems that this structure will be the nucleus
the elimination of turpentine oils during disproportionation at for the design of new BK channel modulators, of interest in the
treatment of acute stroke, epilepsy, asthma, hypertension, gastric
hypermotility, and psychoses (Ohwada et al., 2003). Moreover,
⇑ Corresponding author. Tel.: +34 913934164; fax: +34 913944179. DHA seems to be an anti-inflammatory agent with high PPARa/c
E-mail address: mladero@quim.ucm.es (M. Ladero). dual activation potential and prevent the production of several

0960-8524/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2010.11.022
J.C. Souto et al. / Bioresource Technology 102 (2011) 3504–3511 3505

Nomenclature

CAB abietic acid concentration (mol L1) HPLC high performance liquid chromatography
CD dehydroabietic acid concentration (mol L1) ki kinetic constant
CDi dihydroabietic acid concentration (mol L1) k0 preexponential term of the kinetic constant
CH2 hydrogen concentration (mol L1) MALDI-TOF matrix-assisted laser desorption/ionization
CN neoabietic concentration (mol L1) RAb disappearance rate of abietic acid (mol L1 min1)
CP palustric acid concentration (mol L1) RD appearance rate of dehydroabietic acid (mol L1 min1)
CPi pimaric acid concentration (mol L1) RDi appearance rate of dihydroabietic acid (mol L1 min1)
DHA dehydroabietic acid RN disappearance rate of neoabietic acid (mol L1 min1)
Ea activation energy (J mol1) RPi disappearance rate of pimaric acid (mol L1 min1)
EMA elemental microanalysis T temperature (°C, K)
GCFID gas chromatography with flame ionization detector tR residence time (min)
GC–MS gas chromatography–mass spectrometry Xab abietic acid conversion
H-RMN hydrogen-nuclear magnetic resonance spectroscopy

proinflamatory mediators by activated macrophages and scheme of disproportionation and isomerisation reactions where
differentiated adipocytes. Inasmuch as obesity is associated to several abietic- and pymaric-type rosin acids are considered. The
low-grade inflammation and the production of the mentioned pro- reaction runs were performed on a batch reactor (usual in the
inflamatory mediators, DHA seems to be a compound with interest industry) featuring a double-tier paddle agitator and reaction con-
in the prevention of obesity, and the resistance to insulin associ- ditions were chosen so that kinetic control by chemical reactions
ated to obesity (Kang et al., 2008). This anti-inflammatory activity was assured and oxidation was avoided (by using nitrogen). The
can be related to its capacity to reduce cholesterol and kinetic model, able to fit well all data from 483 to 533 K, is of an
arteriosclerosis. empiric nature, as it is not capable of explaining the high quantity
Disproportionation and isomerisation of rosin acids happen to- of dehydrogenated species compared to hydrogenated species in
gether as rosin is heated (Soltes and Zinkel, 1989). Disproportion- the final product and in samples taken during the kinetic runs.
ation has been considered to be an exchange of hydrogen between Being disproportionation the way to get a rosin of interest
acid molecules, though percentages of DHA higher than 50% are in classic and new applications, and a way of obtaining dehydro-
commonly reached. This reaction has traditionally been catalysed abietic acid of a technical grade, the knowledge of disproportion-
by sulphur alone or with some sulphur organic compounds (as ation kinetics will provide information on the composition of the
sulphides, and disulphides), iron salts, iodide and iodide alkaline reacting mixture under different conditions and time values. An
metals salts, and lithium salts (Soltes and Zinkel, 1989). In the last empirical model can be of use in the design of reactors for a given
decades, due to the toxicity of the gases formed when using sul- catalyst, but a deeper knowledge of the phenomena underlying
phur as catalyst, palladium supported on activated carbon or alu- disproportionation will give more accurate information on the
mina has been used, though inactivation by the iron present in activity of each catalyst, being a sounder basis for the comparison
rosin leads to the necessity of reactivating the catalyst (Matsuo of catalysts. Moreover, the use of commercial catalysts will be of
and Tsuchida, 1981). Nickel and selenium have been tested as cat- interest to industry. Thus, in this work, the disproportionation of
alysts, to replace Pd catalysts, being the search for a non-noble me- rosin with Pd (at 5% concentration) on activated carbon from
tal catalyst one of the most interesting research lines in this field Engelhardt, from 473 to 513 K, is studied, leading to the proposal
(Soltes and Zinkel, 1989). Thus, at the present time, Pd catalysts of a kinetic model based on dehydrogenation, hydrogenation and
show much interest due to the good properties that the dispropor- isomerisation reaction pathway. Runs were performed at the men-
tionated rosin feature (regarding colour, softening-point, low metal tioned temperatures and samples were withdrawn and analyzed
concentration, and resistance to oxidation). by GC-FID and GC–MS. The proposed model and the empirical
Several methods of analysis to determine the composition of ro- model by Wang et al. (2009) are used to fit data obtained with
sin after disproportionation have been used: UV–vis spectropho- the Engelhardt catalyst, using a multiparametric and multivariable
tometry, GC-FID, GC–MS, HPLC, H NMR and MALDI-TOF (Brites non-linear technique with a coupled integration of the ODEs of the
et al., 1993; Gigante et al., 1995; Rigol et al., 2003; Mitani et al., kinetic models.
2007; Kumooka, 2008). UV–vis spectrophotometry, using 254 nm
as wavelength, is a simple method able to give an average conver-
2. Experimental
sion of abietic acid to dehidroabietic acid, and has been the method
of use during a long period of time. Modern chromatographic tech-
2.1. Materials
niques have been applied in the rosin industry since the 80s–90s,
allowing, from that date, to determine the exact composition of
Rosin and glicerol were of technical and pharmaceutical grade,
the disproportinated product. However, little information can be
respectively, and were kindly supplied by LURESA. N,N-Dimethyl-
obtained in the literature regarding the kinetics of the reactions
formamide dimethyl acetal 92% for gas chromatography was pur-
involved, with Pd/C or any other catalyst (Song et al., 1985; Vital
chased from Acros Organics. Pure oleic acid, from Fluka, was
and Lobo, 1992; Wang et al., 2009). Song et al. (1985) used gas
used as internal standard. The catalyst (Escat 111) was 5%Pd on ac-
chromatography to analyse reaction samples from a three-necked
tive carbon and was a kind gift from Engelhard.
round-bottom flask in the disproportionation of gum rosin at
543 K, being dehydroabietic acid the main compound found in
the final samples, with a low quantity of dihydroabietic acids (four 2.2. Catalyst characterisation
of them were identified) and no trace of tetrahydroabietic acids.
Wang et al. (2009) have proposed a kinetic model for the dispro- The specific surface area (BET), total pore volume and average
portionation of rosin on a home-made Pd 5%/C, based on a complex pore diameter were measured in duplicate by nitrogen adsorption
3506 J.C. Souto et al. / Bioresource Technology 102 (2011) 3504–3511

at 77 K in a Coulter SA 3100 Surface Area and Pore Size Analyzer

RAb ¼ k1  C N þ k3  C D  C H2 þ k2  C P  k3  C Ab  k4  C Ab  C H2
(with outgassing of the samples at 573 K during 3 h). The mean
particle size distribution of the catalyst was measured at several
agitation speeds (to avoid aggregation and breakage of the
catalyst) with a Focused Beam Reflectance Measurement with a
LasentecÒ FBRMÒ system model M400L-316K. The morphology of

First order kinetic models (empirical models) and second-order kinetic models (phenomenological models) for the disproportionation of rosin with a Pd 5%/C catalyst used to fit to experimental data in this work.
the catalyst and the EDS was observed in a JEOL JM-6400 scanning
electronic microscope (SEM), using an acceleration voltage of
20 kV. This was also the voltage for the quantitative Electron Dis-
persion Scanning X-ray microanalysis (EDS) controlled by the ISIS
software. Powder X-ray diffraction (XRD) was performed at

 C D  C H2
k = 1.54 nm in a multipurpose PANanalytical diffractometer model
X’Pert MPD with a Cu–K anode, at 40 mV and 200 mA, while scan-
ning the sample over a Bragg angle (2h) from10° to 90°.

k3
0

0
RD ¼ k3  C Ab 

RN ¼ k1  C N
RP ¼ k2  C P
2.3. Esterification runs and sample analysis

Model 5
Kinetic runs were performed between 200 and 240 °C using
0.1% catalyst referred to the rosin mass. Batch runs were started
by charging the reactor with 100 g of rosin into a 250 mL round-
bottom flask with three necks with upper agitation by marine helix
and a distillation head attached designed to condense the tremen-

RAb ¼ k1  C N þ k3  C D  C H2  k3  C Ab  k4  C Ab  C H2
tine. When the reactor was charged a flow of N2 was passed into

RAb ¼ k1  C N þ k2  C P þ k3  C D  k3  C Ab  k4  C Ab
the reactor to avoid oxidation. When reaching 100 °C agitation
was set at 500 r.p.m. At the temperature of reaction, a zero time
sample was withdrawn and catalyst was added, taking more sam-
ples during 6 h.

 C D  C H2
Aliquots of the samples have also been crushed to powder to
analyze their percentual mass content in C and H by elemental
microanalysis (EMA) using each time a 1 mg aliquot and a Perkin

k3
0

0
RD ¼ k3  C Ab þ k2  C P 
Elmer CHN 2400 elemental microanalyser.
RD ¼ k3  C Ab  k3  C D
For a qualitative and quantitative analysis of the rosin acids in-

0
0

volved in the isomerisation and disproportionation reaction net-

RN ¼ k1  C N
RP ¼ k2  C P
work, a GC–MS analytical procedure was also employed. A HP
6890 GC–MS chromatograph equipped with a mass detector MSD
Model 2

Model 4

RH2 ¼ k3  C D  C H2  k4  C Ab  C H2  k3  C D  C H2  k5  C Pi  C H2  k6  C H2
type 5973 was used to analyze the reaction compounds. The sam-
ple constituents were separated with a HP-INNOWAX (crosslinked
PEG) 30 m  0.32 mm ØI  0.25 lm column. Helium (purity
P99.999%) with a flow-rate of 16 mL min1, and a constant pres-
sure of 24 kPa was used as the carrier gas. The oven temperature
was T0 = 150 °C, 1 min, rate 5 °C/min, T1 = 190 °C, rate 10 °C/min,
RAb ¼ k1  C P þ k3  C D  C H2  k3  C Ab  k4  C Ab  C H2

T2 = 240 °C, 25 min, and post run, 249 °C, 5 min. The injection port
temperature was 210 °C, and the detector temperature was 250 °C.
The injection volume was 1 ll (splitless mode). Data acquisition
RAb ¼ k1  C N þ k2  C P  k3  C Ab  k4  C Ab

was performed in scan mode. Mass spectra were recorded from


0
 C D  C H2

m/z 50–550 amu.


To determine the concentration of the main rosin acids, GC-
Second-order kinetic models

FID analyses of withdrawn samples were performed. Samples


First-order kinetic models

k3
0

were methylated with N,N-dimethylformamiddimethylacetal and


RD ¼ k3  C Ab þ k2  C N 

analyzed on a HP model 5890 series II gas cromatograph with a


RPi ¼ k5  C Pi  C H2
RDi ¼ k4  C Ab  C H2

flame ionization detector using a dimethylpolysiloxane capillary


0

column (15 m  0.32 mm  0.5 mm, Zebron 1, Phenomenex).


RPi ¼ k5  C Pi
RN ¼ k1  C N

RN ¼ k2  C N
RDi ¼ k4  C Ab
RP ¼ k2  C P

RP ¼ k1  C P
RD ¼ k3  C Ab

The operating conditions of the gas chromatography were as fol-


Model 1

Model 3

lows: (i) initial oven temperature 453 K for 0 min, program at


1 K min1 to 468 K, then at 20 K min1 to 558 K and hold for
5.5 min, with total run time of 25 min; (ii) N2 was used as carried
gas with GC-FID system; (iii) the injection temperature was 523 K
and the detectors temperature was 573 K (Rigol et al., 2003; HSE,
2006).
Dihydroabietic acids

Dihydroabietic acids
Dehydroabietic acid

Dehydroabietic acid

2.4. Kinetic modelling and simulation


Neoabietic acid

Neoabietic acid
Palustric acid

Palustric acid
Pimaric acid

Pimaric acid
Abietic acid

Abietic acid
Compound

Hydrogen

Two kinetic models were the base of the proposed models to be


fitted to data from the reaction system: a first-order kinetic model
Table 1

(Wang et al., 2009) and a second-order kinetic model that consid-


ers dehydrogenation of abietic-type rosin acids followed by hydro-
J.C. Souto et al. / Bioresource Technology 102 (2011) 3504–3511 3507

genation of abietic- and pimaric-type rosin acids with the liberated 3.2. Reaction route and kinetic modelling
hydrogen from the previous reactions. In the first-order kinetic
model, it was assumed that the reactions just involve isomeriza- To fit kinetic models to molar quantities of the involved
tion and exchange of hydrogen among molecules of rosin resin rosin acids a non-linear regression technique based on the
acids, a double bond rearrangement of rosin resin acids molecules Marquardt–Levenberg algorithm has been applied. Several kinetic
(Wang et al., 2009). In the second-order kinetic model, the concen- models have been considered, based on the isomerisation
tration of hydrogen able to perform hydrogenation of pimaric and pathways in the literature as well as assumptions and facts in
abietic acids was considered. The proposed kinetic models in- references regarding disproportionation. These kinetic models are
cluded model 1 (the one by Wang et al.), the same with equilib- shown on Table 1, based on an empirical first order model (models
rium between abietic and dehydroabietic acid (model 2) and 1 and 2) and on second-order kinetics, of a more phenomenological
three kinetic models based on second-order kinetic laws and the nature (models 3–5). Identified compounds by GC–MS comprised
existence of free hydrogen molecules (models 3–5). Model 3 in- abietic-type acids (abietic, palustric, neoabietic, dihydroabietic
volves that palustric acid isomerised to abietic acid, while neoabi- and dehydroabietic acids) as well as several pimaric and dihydro-
etic acid is directly dehydrogenated. Model 4 considers exactly the pimaric acids (several peaks for each one, as the inner double bond
opposite: neoabietic acid isomerise to abietic acid, while palustric can be in several positions inside the cyclic structure). Chromato-
acid is converted to dehydroabietic acid. Finally, model 5 involves graph profile was identical in GC-FID and GC–MS, so peaks were
the isomerisation of both acids to abietic acid prior to the dehydro- assigned in a straightforward way, though during the construction
genation of the latter. Equations of all kinetic models are included of the corresponding calibrates in GC-FID with several rosin acid
in Table 1. standards, peak assignations were confirmed.
Kinetic models were fitted using the Marquardt–Levenberg Disproportionation of rosin involves dehydrogenation of abi-
algorithm together with a Runge–Kutta method for the numerical etic-type acids, as well as hydrogenation and isomerisation reac-
integration of the kinetic equations. The selection of the most tions, so that the mixture of acids evolves to a final composition
appropriated model was based on the usual physical criteria (posi- that is more stable from a thermodynamical viewpoint. It is pro-
tive value of the kinetic parameters and values for the activation posed that abietic acid is dehydrogenated to dehydroabietic acid
energies within adequate ranges) and statistical criteria (Student’s (which is the actual sought after product in industry), so that
t value for each kinetic parameter, Fischer’s F and SQR values for two conjugated double bonds turn into an aromatic ring. At the
each kinetic model). same time, palustric and neoabietic acid convert to abietic acid
or one of them can be dehydrogenated. Pymaric type acids (several
of which are identified in the GC–MS chromatograms) are hydroge-
3. Results and discussion nated to dihydropymaric acids. The double bond external to the
cyclic structure typical of such acids is prone to hydrogenation.
3.1. Catalyst characterisation The hydrogen molecules from the dehydrogenation step can re-
act with pymaric and abietic-type acids, or can escape from the li-
Basic characteristics of the Escat 111 catalyst were obtained in quid phase where all reactions happen. Although this latter fact is
the Coulter SA 3100 Analyser. BET surface area was 582 m2/g, more than possible, according to results from all authors (Song
while total pore volume was 0.50 ml/g. BET surface area related et al., 1985; Wang et al., 2009), it is also thought that hydrogen
to micropores (less than 5 nm diameter) was 200 m2/g, with an in- atoms are rearranged among the different acids involved so that
ner volume due to micropores of 0.088 ml/g. The pore size distri- dehydrogenation and hydrogenation are synchronized and no
bution shows a high percentage of volume related to the hydrogen leaves the liquid phase and, even, a very recent kinetic
micropore region, while macropores in the 20–80 nm pore diame- study seems to be based on such assumption (Wang et al., 2009).
ter region are present (data corroborated by SEM at 500–15,000 The equilibrium between abietic and dehydroabietic acids
enlargement). Mesopores are also of importance, with a 35% of shows that the former is really stable at the high temperatures
the total volume linked to pores in the 2–20 nm pore diameter re- where disproportionation and other reactions involving rosin take
gion. Thus, pore distribution is almost trimodal, mainly bimodal, place. In fact, at 200 °C, more than 80% of the rosin acids not yet
but highly dispersed. dehydrogenated is abietic acid (in the absence of a disproportion-
XRD shows little crystallinity but this is linked to the presence ation catalyst), so it is the most stable acid among the abietates.
of Pd crystals (as SEM semiquantitative elemental analysis con- Moreover, a worse fitting and an even worse value of the F param-
firms). It can be seen that the sample exhibited four peaks at 2h eter are obtained when using a non-equilibrium considering ki-
of 34.04°, 43.53°, 55.05° and 80.01°, ascribed, respectively, to netic model to fit the experimental data herein presented as well
(1 1 1), (2 0 0), (2 1 0) and (3 1 1) reflections of Pd metal with a face as the data featured in the paper by Wang et al. (2009), as shown
centered cubic (fcc) structure. The average size of the Pd crystallite in Table 2.
particles was estimated from the full width at half maximum of the If the empirical model involving only the redistribution of
diffraction peaks to be 2.92 nm through Scherrer equation. The dis- hydrogen atoms among the rosin acid molecules is modified in
persion values (D), the ratio of number of surface atoms to the total such a way that there is a chemical equilibrium between abietic
number of atoms is calculated from the equation D = 1.13/d, where and dehydroabietic acid, a model that fits considerably better to
d is the crystallite size (2.9 nm). Thus, %D = 1.13/2.92  100 = 38.6% experimental data is obtained. (F changes from 1090 to 2649
(Pattabiraman, 1997). although two kinetic parameters have to be added to those in mod-
Particle size is small, as the FBRM measurements showed, with el 1, and SQR is reduced by a factor of 2.5, as given in Table 2). All
an average diameter of 1.6 ± 1.4 nm (high dispersion) and a totally these results seem to show that not only abietic acid and pimaric
asymmetric distribution, with a maximum at 0.48 nm. This is con- acid react with free hydrogen molecules, but also that dehydroabi-
firmed by SEM images, with a high number of small particles in the etic acid itself is prone to be hydrogenated.
micro-meter range and some big particles with hydraulic diame- Fitting of the proposed kinetic models to experimental data has
ters up to 100 lm. EDS measurements give an idea of the distribu- been performed by first calculating parameters at given tempera-
tion of Pd on carbon particles, with a percentage of Pd between tures. Thus, values of the parameters at each temperature have
2.9% and 4.3% depending on the particle and zone of particle been calculated for each model and, afterwards, by linearization
studied. of the Arrhenius equation, values for the neperian logarithm of
3508 J.C. Souto et al. / Bioresource Technology 102 (2011) 3504–3511

Table 2
Fit of the empirical and phenomenological models to experimental data of disproportionation of rosin obtained at several temperatures (from 200 to 240 °C) with an industrial Pd
5%/C catalyst from Engelhardt.

Kinetic parameters Model 1 Model 2 Model 3 Model 4 Model 5


Valor ± STD error Valor ± STD error Valor ± STD error Valor ± STD error Valor ± STD error
k1 Ln k0 21.8 ± 12.21 21.9 ± 8.08 19.4 ± 2.86 13.9 ± 2.43 18.8 ± 2.76
Ea/R 12,221 ± 5768 12,225 ± 3879 11,005 ± 1379 8450 ± 1175 10,797 ± 1333
k2 Ln k0 18.0 ± 5.13 18.5 ± 3.35 13.2 ± 5.21 22.1 ± 8.03 23.0 ± 6.80
Ea/R 10,422 ± 2481 10,635 ± 1619 8066 ± 2509 12,166 ± 3847 12,761 ± 3265
k3 Ln k0 18.7 ± 1.15 18.4 ± 0.70 21.1 ± 0.85 22.2 ± 1.03 19.2 ± 0.76
Ea/R 11,116 ± 558 10,876 ± 340 12,276 ± 418 12,845 ± 504 11,221 ± 371
0
k3 Ln k0 – 5.6 ± 2.60 15.7 ± 8.01 23.2 ± 7.00 18.8 ± 6.83
Ea/R – 5901 ± 1301 8587 ± 3945 12,399 ± 3408 10,361 ± 3342
k4 Ln k0 8.2 ± 2.27 14.0 ± 1.23 17.9 ± 4.89 25.1 ± 4.10 20.0 ± 3.49
Ea/R 6673 ± 1109 9407 ± 602 9472 ± 2392 13,195 ± 1994 10,711 ± 1707
k5 Ln k0 22.3 ± 10.62 22.3 ± 6.59 13.5 ± 8.63 18.8 ± 6.61 14.5 ± 6.36
Ea/R 13,434 ± 5244 13,434 ± 3253 8606 ± 4245 11,066 ± 3235 9041 ± 3129
k6 Ln k0 – – 23.9 ± 7.63 32.1 ± 6.77 28.1 ± 6.33
Ea/R – – 12,327 ± 3751 16,534 ± 3327 14,637 ± 3118
Model statistical F 1090 2649 3604 3292 3724
parameters SQR 6.62  102 2.53  102 1.63  102 1.77  102 1.59  102

Note: k1–k6 in L mol1 min1. Ea/R (K).

the preexponential parameters and for the activation energies have 1992). Thus, model 4 can be considered not valid based on
been obtained. Finally, fitting of the kinetic models to all experi- literature and on the statistical analysis in Table 2. Models 3 and
mental data (at the several temperature values used) has been per- 5 are both statistically valid, and there are reasons in the literature
formed to calculate the optimal values of the kinetic parameters. for both of the chemical routes to be realistic. Model 5 involves the
As shown in Table 2, the empirical model based on first-order ki- rearrangement of two double bonds in neoabietic acid to render
netic equations is able to reasonably fit experimental data (F value abietic acid. This isomerisation is supported by some references
higher than the critical one for the given values of data and param- in literature (Takeda et al., 1969; Perelson et al., 1990).
eter numbers), but more complex models fit much better experi- This work and the one by Wang et al. (2009) show that abietic
mental data, reducing the SQR value by a factor of 4 and acid is the main acid of gum rosin and converts directly to dehy-
increasing by the same factor the F value when compared to model droabietic acid. The statistical study in this work shows that an
1. Thus, both from a statistical and from a physical point of view, equilibrium exists between abietic and dehydroabietic acid. Some
disproportionation of rosin proceeds through the dehydrogenation authors suggest that levopimaric acid and palustric acid can be
of rosin acids of the abietate kind, and the hydrogenation of all considered as the intermediates between abietic acid and dehydro-
acids involved, affecting mainly to pimaric acids, as the hydrogena- abietic acid (Enoki, 1976; Portugal et al., 1996), and this seems to
tion of the exocyclic double bonds is favoured both by thermody- be the case when temperature is lower than 200 °C. The common
namics and kinetics (Song et al., 1985; Wang et al., 2009). As industrial practice is, however, to work in the range 200–270 °C,
hydrogen molecules diffused fast, a great percentage of this gas depending on the catalyst used. Here, abietic acid and dehydroabi-
is lost from the reacting mass. etic acid seem to be the most stable acids. In fact, when using Pd on
While palustric acid can be dehydrogenated directly, there is no charcoal as catalyst during hydrogenation of abietic acid, the cata-
evidence that it cannot be converted to abietic acid previously lyst acts as a hydrogenation catalyst at temperatures lower than
(model 5, SQR = 0.0159, F = 3724). In fact, while model 4 has a 150 °C and as a dehydrogenation catalyst at higher temperatures
SQR value slightly worse than models 3 and 5 (SQR = 0.0177, (Yu et al., 2005).
F = 3292), model 3 is statistically equal to model 5 (SQR = 0.0163; Thus, considering the information in the literature and the re-
F = 3604). Considering the work on isomerisation of levopimaric sults herein presented, model 5 seems to be more reasonable than
acid by Vital and Lobo (1992), this acid reacted directly to neoabi- model 3 (neoabietic and palustric acids react to abietic acid prior to
etic, abietic and palustric acids, being these two latter acids related the dehydrogenation/hydrogenation steps), and dehydrogenation
by a chemical balance. Vital and Lobo (1992) suggest that neoabi- is the most important reaction due to the high operational temper-
etic acid is not related directly to abietic acid, but palustric acid atures. In Fig. 1, experimental data are shown as points, while the
converts to this latter acid, supporting model 3. In this work and fitting of model 5 to them is drawn in lines. Moreover, according to
another from some years later (Portugal et al., 1996), results sug- kinetic model 5 and as shown in Fig. 2, hydrogen molar amount
gest that levopimaric acid is stable only at relatively low tempera- grows initially until a maximum is reached and, afterwards, grad-
tures (those of rosin distillation, 150–180 °C), and isomerised ually decreases due to diffusion to the atmosphere, and this pro-
directly to abietic, neoabietic and palustric acid. In fact, levopima- cess is faster the higher temperature is.
ric acid is the most abundant rosin acid in fresh oleoresin, while All references in literature regarding disproportionation of ro-
abietic acid is the prevalent in heat- or acid-treated oleoresin sin, as well as the results from this work, lead to the proposal of
(Enoki, 1976). On the other hand, abietic acid isomerises to both simple potential kinetic models, though many reactions are in-
palustric and neoabietic acids, with a higher percentage of the volved in the final reaction scheme. Considering that reactions take
former, remaining the percentage of abietic acid higher than 80% place on or near the catalyst surface, hyperbolic kinetic equations
at temperatures of 200 °C or higher (Takeda et al., 1969). Palustric could be expected (for example, regarding the adsorption and
acid isomerisation leads to abietic acid, the prevalent, and neoabi- splitting of hydrogen molecules or the adsorption of rosin acid
etic acid (Joye and Lawrence, 1961), though it is suggested that molecules). The macroscopic kinetic model shows, however, that
neoabietic acid comes from levopimaric acid, being a product of these surface phenomena are too fast to be considered in the
isomerisation of palustric acid by a rearrangement of double bonds kinetic equations. As sulphur is the homogeneous catalyst with
in one of the cycles of the diterpenic structure (Vital and Lobo, an activity similar to that of Pd/C, and its activity is related to its
J.C. Souto et al. / Bioresource Technology 102 (2011) 3504–3511 3509

0.30 0.35

(a) (b)
0.30
0.25

0.25
molar quantity (mol)

0.20

molar quantity (mol)


0.20
0.15
0.15

0.10
0.10

0.05
0.05

0.00 0.00

0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time (min) Time (min)

0.35 0.35

(c) (d)
0.30 0.30

0.25 0.25
molar quantity (mol)
molar quantity (mol)

0.20 0.20

0.15 0.15

0.10 0.10

0.05 0.05

0.00 0.00

0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time (min) Time (min)

0.35

(e)
0.30

0.25
molar quantity (mol)

0.20

0.15

0.10

0.05

0.00

0 50 100 150 200


Time (min)

Fig. 1. The phenomenological kinetic model fitting to experimental data obtained at (a) 473 K; (b) 483 K; (c) 493 K; (d) 503 K and (e) 513 K. (j) Abietic acid, (N)
dehydroabietic acid, (}) dihydroabietic acid, (s) palustric acid, (H) neoabietic acid, (h) pimaric acid, (–) kinetic model.

ability to withdraw hydrogen atoms and transfer them to other products is taking place near the surface, so the system is acting
molecules, it can be envisaged a similar behaviour for the Pd atoms as a pseudo-homogeneous reacting system (Kanno et al., 1987;
at the charcoal surface. There are a number of reasons why a mac- Zhang et al., 2010).
roscopic kinetic model has simply first-order kinetic equations When comparing kinetic parameters and model to those ob-
even if heterogeneous catalysis is involved. However, all of them tained by Wang et al. (2009), it can be observed that values for acti-
show that only one phenomenon is controlling the overall rate of vation energies lay in the range 80–100 kJ mol1, regardless of the
the process and, what is more, no accumulation of reactants and results considered (ours or those of Wang and coworkers). This
3510 J.C. Souto et al. / Bioresource Technology 102 (2011) 3504–3511

GC–MS not only is employed for the identification of the com-


0.06 pounds in the sample, depending on the value of a quality factor,
473 K but it can also be used to quantify the amount of each compound.
0.05
Here again, dehydrogenation seems to happen, as the mass of the
dehydrogenated species is considerably higher than the sum of
molar amount (mol)

the several dehydrogenated species that have been identified, even


0.04 483 K if both pymaric and abietate structures are taken into account.
For an accurate identification of the several acids involved in
0.03 the reactions taking place, GC-FID has been used, using standards
of the most important rosin acid identified by CC–MS and oleic acid
493 K
0.02 as an internal standard. Through this analytical strategy, quantifi-
cation of the rosin acids concentration in the samples is performed.
503 K By adding amounts of dehydrogenated species on one hand, and
0.01
performing the same calculation for the dehydrogenated ones on
513 K the other, it can be seen that the former is almost three times lar-
0.00
ger than the latter, and this is so independently of the reaction
0 50 100 150 200 250
temperature. Though not so clearly, the same trend is observed
Time (min)
in the paper of Wang and coworkers (2009). This fact is also due
Fig. 2. Simulated hydrogen molar amount temporal evolution in the reaction liquid to the statistically identical values of activation energies in dehy-
using kinetic model 5. drogenation and hydrogenation reactions (Wang et al., 2009). Thus,
dehydrogenation is the prevailing reaction in the reaction network
taking place during rosin disproportionation, and this reaction
could be expected as results from Wang and coworkers and our re-
leads to the most stable rosin acid to oxidative degradation: dehy-
sults are obtained in a similar range of particle size (less than
droabietic acid.
60 mm in diameter) and agitation speed (more than 400 rpm),
Palladium on charcoal catalysts seems to behave as sulphur,
where no hindrance due to mass transfer phenomena are expected
highly increasing disproportionation rates at relatively low tem-
and chemical reactions are the controlling step.
peratures (comparing to other rosin processes, as esterification,
As said previously, despite the empirical model (model 1) not
which proceeds at 260–290 °C). This could be so as Pd atoms act
being the best one, it is able to be fit to all empirical data reason-
as a hydrogen adsorbent, reservoir and a hydrogen-transfer vehi-
ably well: residual sum of squares is low, F value is high and error
cle, while the main reactions take place in the liquid phase, which
ranges for the parameters are narrow. To compare results of both
is a mixture of rosin acids (it is a solvent-less reacting system). In
works, it is necessary to take into account the fact that rosin and
fact, the formation of palladium hydrides and the well-known dis-
catalyst concentration in Wang’s work are 55% the ones used in
ruptive chemisorption of hydrogen on the palladium surface could
this work, as here solvent-less conditions are employed, while
be part of the explanation why there is a higher amount of dehy-
Wang and coworkers dilute rosin by using a mineral oil (oil no.
drogenated than hydrogenated rosin acids (Markus et al., 2007;
200) in a volume ratio 45:55 to rosin. Moreover, catalysts are sim-
Tomaszewska et al., 2008; Weng et al., 2010). On the other hand,
ilar but not identical (crystallinity and dispersion of palladium are
hydrogen escapes from the reacting liquid aided by the nitrogen
different). In this situation, kinetic constants of isomerisation reac-
that flows onto and over the liquid surface, through an irreversible
tions (k1 and k2), the kinetic constant value for the dehydrogena-
process that follows first-order kinetics and that highly depends on
tion reaction (k3), and the kinetic constant for pimaric acids
temperature, as the high activation energy of constant k6 of model
hydrogenation (k5) are similar in both works, while the hydrogena-
5 shows. In a similar way, but involving a chemical reaction, sul-
tion reaction constant for dehydroabietic acid (k4) is eight times
phur withdraws hydrogen atoms from rosin acids yielding hydro-
lower in this work. Thus, conditions in this paper are favourable
gen sulphide, that also desorbs partly from the liquid (Zhao et al.,
to isomerisation and dehydrogenation of abietic acids and less
2008).
favourable for hydrogenation reactions (mostly in the case of abi-
etic acids). In Wang’s paper, the amount of hydrogenated acids
is, more or less, between 60% and 75% that of dehydroabietic acid. 4. Conclusions
Here, hydrogenated acids amount to less than 40% of the dehydro-
genated acid, so, on the whole, both catalysts used are favourable In this paper, the role of hydrogen in rosin disproportionation is
to dehydrogenation. clarified with the aid of recent literature and using several analyt-
ical techniques. Dehydrogenation of abietic acid extent is almost
three times higher than that of hydrogenation reactions pooled
3.3. Hydrogen fate during disproportionation
all together, so, in the presence of Pd/C as a catalyst, dehydrogena-
tion is the prevailing phenomenon.
Two approaches have been considered to prove the role of
Based on this fact, a phenomenological kinetic model is se-
hydrogen in the disproportionation of rosin: elemental microanal-
lected, after discrimination among several kinetic models and reac-
ysis (EMA) and gas chromatography, using both a quadrupole mass
tion schemes was performed by non-linear regression techniques.
detector and a flame ionization detector (FID).
This model considers isomerisation reactions of abietates, as well
The hydrogen to carbon mass ratio percentage can be calculated
as dehydrogenation and hydrogenation of abietic acid and hydro-
from data obtained by EMA and it changes from 1.25 to 1.17 from
genation of pimaric acids.
pure abietic to pure dehydroabietic acid. As the runs at all the tem-
peratures proceed, this ratio changed from 1.22 to 1.19–1.18 in all
cases, showing that certain dehydrogenation of the solid material Acknowledgements
happens in every run. However, as the material is not pure and
some turpentine distilled off while disproportionation proceeds, Financial support from the Spanish Ministry of Education and
this technique allows only for a qualitative assessment of the from L.U.R.E.S.A. (through project PETRI 95-0821.OP) is gratefully
dehydrogenation. acknowledged. Moreover, the authors want to extent this
J.C. Souto et al. / Bioresource Technology 102 (2011) 3504–3511 3511

recognition to the Spanish Ministry of Environment (project MMA- Mayer, M.J.J., Meuldijk, J., Thoenes, D., 1996. Emulsion polymerization of styrene
with disproportionated rosin acid soap as emulsifier. Journal of Applied
0392063-11.2) for further financial support. The authors also want
Polymer Science 59 (6), 1047.
to express their deepest gratitude to Engelhardt Inc. (now, branch Mitani, K., Fujioka, M., Uchida, A., Kataoka, H., 2007. Analysis of abietic acid and
of BASF) for the kind gift of a 5%Pd/C catalyst, as well as to L.U.R.- dehydroabietic acid in food samples by in-tube solid-phase microextraction
E.S.A. for the generous gift of industrial rosin. coupled with liquid chromatography–mass spectrometry. Journal of
Chromatography A 1146 (1), 61–66.
Ohwada, T., Nonomura, T., Maki, K., Sakamoto, K., Ohya, S., Muraki, K., Imaizumi, Y.,
References 2003. Dehydroabietic acid derivatives as a novel scaffold for large-conductance
calcium-activated K+ channel openers. Bioorganic and Medicinal Chemistry
Brites, M.J., Guerreiro, A., Gigante, B., Marcelo-Curto, M.J., 1993. Quantitative Letters 13 (22), 3971–3974.
determination of dehydroabietic acid methyl ester in disproportionated rosin. Pattabiraman, R., 1997. Electrochemical investigations on carbon supported
Journal of Chromatography 641 (1), 199–202. palladium catalysts. Applied Catalysis A: General 153 (1–2), 9–20.
Duan, W., Shen, C., Fang, H., Li, G.H., 2009. Synthesis of dehydroabietic acid- Perelson, M.E., Dmitrieva, L.K., Vodolazskaya, V.M., Semichev, V.P., 1990.
modified chitosan and its drug release behavior. Carbohydrate Research 344 (1), Quantitative determination of the sum of the acids of the abietic type in
9–13. rosin. Chemistry of Natural Compounds 26 (4), 390–392.
Enoki, A., 1976. Isomerization and autoxidation of resin acids. Wood Research: Portugal, I., Vital, J., Lobo, L.S., 1996. Isomerization of resin acids during pine
Bulletin of the Wood Research Institute Kyoto University 59 (60), 49–57. oleoresin distillation. Chemical Engineering Science 51 (11), 2577–2582.
Fonseca, T., Gigante, B., Marques, M.M., Gilchrist, T.L., De Clercq, E., 2004. Synthesis Rigol, A., Latorre, A., Lacorte, S., Barceló, D., 2003. Direct determination of resin and
and antiviral evaluation of benzimidazoles, quinoxalines and indoles from fatty acids in process waters of paper industries by liquid chromatography/
dehydroabietic acid. Bioorganic & Medicinal Chemistry 12 (1), 103–112. mass spectrometry. Journal of Mass Spectrometry 38 (4), 417–426.
Gigante, B., Santos, L., Marcelo-Curto, M.J., Ascenso, J., 1995. 13C and 1H NMR Soltes, E.J., Zinkel, D.F., 1989. Chemistry of rosin. In: Zinkel, D.F., Russell, J. (Eds.),
assignments for a series of dehydroabietic acid derivatives. Magnetic Resonance Naval stores: Production, Chemistry, and Utilization. Pulp Chemical Association,
in Chemistry 33 (4), 318–321. New York.
González, M.A., Pérez-Guaita, D., Correa-Royero, J., Zapata, B., Agudelo, L., Mesa- Song, Z.-Q., Zavarin, E., Zinkel, D.F., 1985. On the palladium-on-charcoal
Arango, A., Betancur-Galvis, L., 2010. Synthesis and biological evaluation of disproportionation of rosin. Journal of Wood Chemistry and Technology 5 (4),
dehydroabietic acid derivatives. European Journal of Medicinal Chemistry 45 535–542.
(2), 811–816. Takeda, H., Schuller, W.H., Lawrence, R.V., 1969. Thermal behavior of some resin
HSE, 2006. MDHS 83/3 resin acids in rosin (colophony) solder flux fume. Available acid esters. Journal of Chemical and Engineering Data 14 (1), 89–90.
from: <www.hse.gov.uk/pubns/mdhs/pdfs/mdhs83-2.pdf>. Tanaka, R., Tokuda, H., Ezaki, Y., 2008. Cancer chemopreventive activity of rosin
Hu, L.-h., Zhou, Y.-h., Song, Z.-q., 2006. Synthesis and properties of rosin-based constituents of Pinus spez. and their derivatives in two-stage mouse skin
polyglucoside. Linchan Huaxue Yu Gongye 26 (1), 11–14. carcinogenesis test. Phytomedicine 15 (11), 985–992.
Joye, N.M., Lawrence, R.V., 1961. The thermal isomerization of palustric acid. The Tomaszewska, A., Ciszewski, A., Stepien, Z.M., 2008. Interaction of hydrogen with
Journal of Organic Chemistry 26 (4), 1024–1026. palladium surface. FIM and FEM studies. Applied Surface Science 254 (14),
Kang, M.-S., Hirai, S., Goto, T., Kuroyanagi, K., Lee, J.-Y., Uemura, T., Ezaki, Y., 4386–4390.
Takahashi, N., Kawada, T., 2008. Dehydroabietic acid, a phytochemical, acts as Vital, I.P.J., Lobo, L.S., 1992. Resin acids isomerization: a kinetic study. Chemical
ligand for PPARs in macrophages and adipocytes to regulate inflammation. Engineering Science 47 (9–11), 2671–2676.
Biochemical and Biophysical Research Communications 369 (2), 333–338. Wang, L., Chen, X., Liang, J., Chen, Y., Pu, X., Tong, Z., 2009. Kinetics of the catalytic
Kanno, T., Kimura, T., Onose, T., Hayashi, M., Kobayashi, M., 1987. Evaluation of the isomerization and disproportionation of rosin over carbon-supported
first order approximation in heterogeneous catalysis. Memoirs of the Kitami palladium. Engineering Journal 152 (1), 242–250.
Institute of Technology 19 (1), 71–88. Weng, B.C., Yu, X.B., Wu, Z., Li, Z.L., Huang, T.S., Xu, N.X., Ni, J., 2010. Improved
Kumooka, Y., 2008. Analysis of rosin and modified rosin esters in adhesives by dehydrogenation performance of LiBH4/MgH2 composite with Pd nanoparticles
matrix-assisted laser desorption/ionization time-of-flight. Forensic Science addition. Journal of Alloys and Compounds 503 (2), 345–349.
International 176 (2), 111–120. Yu, S.-t. L., L., Liu, F.-s., Zhang, S.-f., Yang, J.-z., 2005. Studies on hydrogenation of
Markus, H., Plomp, A.J., Sandberg, T., Nieminen, V., Bitter, J.H., Murzin, D.Y., 2007. rosin over Pd/C in supercritical CO2. The Proceedings of the Third International
Dehydrogenation of hydroxymatairesinol to oxomatairesinol over carbon Conference on Functional Molecules, pp. 234–237.
nanofibre-supported palladium catalysts. Journal of Molecular Catalysis A: Zhang, L., Sheng, B., Xin, Z., Liu, Q., Sun, S., 2010. Kinetics of transesterification of
Chemical 274 (1–2), 42–49. palm oil and dimethyl carbonate for biodiesel production at the catalysis of
Matsuo, K., Tsuchida, S. 1981. Stabilized Rosin Ester and Pressure-Sensitive heterogeneous base catalyst. Bioresource Technology 101 (21), 8144–8150.
Adhesive and Hot-Melt Composition Based Thereon – US Patent Application: Zhao, G.B.R., 2008. LA, US, Cooke, Todd Maxwell (Greenwell Springs, LA, US).
06/169,619. Patent Number: 4302371. Method of Producing Disproportionated Rosin. Albemarle Corporation, Baton
Mayer, M.J.J., Meuldijk, J., Thoenes, D., 1995. Influence of disproportionated rosin Rouge, LA, US, United States.
acid soap on the emulsion polymerization kinetics of styrene. Journal of Applied
Polymer Science 56 (2), 119–126.

You might also like