Vous êtes sur la page 1sur 8

View Article Online / Journal Homepage / Table of Contents for this issue

PAPER

www.rsc.org/pccp | Physical Chemistry Chemical Physics

Chemorheological analysis and model-free kinetics of acid catalysed


furfuryl alcohol polymerization

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

Nathanael Guigo, Alice Mija,* Luc Vincent and Nicolas Sbirrazzuoli


Received 25th May 2007, Accepted 24th July 2007
First published as an Advance Article on the web 8th August 2007
DOI: 10.1039/b707950h
The complete curing of furfuryl alcohol (FA), was studied by chemorheological analysis and
model-free kinetics under isothermal and non-isothermal modes. Polymerization of FA under
acidic catalysis involves complex reactions, with several steps (such as condensations and
DielsAlder cycloadditions). To account for the polymerization complexity, kinetic analysis of
DSC data was performed with a model-free isoconversional method. The obtained
Ea-dependencies were closely-correlated with the variation of complex viscosity during curing.
Linear condensations are predominant during the early curing stage and are followed by two
distinct stages of branching cycloadditions. Gelation and vitrication, identied by rheometric
measurements, were associated with a decrease of the overall reaction rate that becomes
controlled by diusion of small oligomers. Before vitrication, the rate of crosslinking is limited
by the mobility of longer polymer chains and diusion encounters a large energy barrier due to
the cooperative nature of the motions, leading to higher Ea values.

Introduction
Vegetable biomass represents a sustainable solution to replace
petroleum-based chemicals. In particular, hemicellulose constitutes an important source of monomers such as furfuryl
alcohol (FA). This latter is obtained by conversion of furfural
and can react with Brnsted or Lewis acidic catalysts to form a
black, cross-linked macromolecule.1 At an industrial scale, FA
is usually converted into prepolymers and these furanic resins
represent excellent eco-friendly precursors for wood impregnation2 and for elaborating composite materials3,4 or wood
adhesives.5 Currently, for this type of resin, the control of
prepolymerization is an important issue and requires careful
application of processing conditions. A perfect knowledge of
the chemorheological behaviour of the reactive polymer system during processing is very important in order to determine
optimum process parameters, adapted to the aimed application. As it is the case for many thermosets, viscosity of FA
prepolymers is strongly dependent on the evolution of temperature and the extent of conversion during curing. The
variation of the viscosity is also a key parameter which may
govern the chemical reactions at the microscopic scale.
The polymerization reaction of furfuryl alcohol has been
previously studied under dierent experimental conditions.613
The reported step-growth curing mechanisms can be separated
into two stages. In the rst stage, under acid catalysis, the
methylol group of one furan ring condenses with the C5
position of another furan ring with dehydration.1 Furan rings
connected by methylene linkages create linear oligomers,14 as
shown in Scheme 1. In the second stage, these linear oligomers

Thermokinetic Group, Laboratory of Chemistry of Organic and


Metallic Materials C.M.O.M., Institute of Chemistry of Nice,
University of NiceSophia Antipolis, 06108 Nice Cedex 2, France.
E-mail: mija@unice.fr

This journal is


c

the Owner Societies 2007

are cross-linked into black materials. It is postulated3,6,10,15,16


that these oligomers are branched together mostly due to
DielsAlder cycloadditions between the furan rings (diene)
and the dihydrofuranic cycles (dienophile). Despite these
previous investigations, the overall kinetics of this complex
curing process (i.e., linear growth of chains and crosslinking)
still remain imperfectly understood. According to Milkovic,17
the overall mechanism involves many steps that are likely to
have dierent activation energies. The contribution of these
steps into the overall cure rate should generally vary with both
temperature and extent of curing. This means that the eective
activation energy determined from the overall rate measurements is likely to be a function of these two variables. A kinetic
study based on empirical models could be ineective if the
reaction models are unknown. For this reason, an alternative
solution is to use model-free isoconversional methods. These
methods require no hypotheses on the reaction mechanism
and allow for evaluation of the apparent activation energy as a
function of the extent of conversion. Consequently, changes in
curing mechanisms are associated with the variation of apparent activation energy.
The objective of the present study is to get a better understanding of the curing behaviour of FA and to highlight
changes in mechanisms during both resinication and crosslinking. For this purpose, the curing of FA catalyzed with
maleic anhydride (MA) was investigated by infrared spectroscopy (IR), rheometry and by dierential scanning calorimetry
(DSC) under both isothermal and non-isothermal conditions.
DSC data were treated with an advanced isoconversional
method in order to yield the dependence of activation energy
on conversion. To our knowledge, this work reports the rst
chemorheological study combining model-free kinetics and
rheological data for the study of FA polymerization. In
particular, we demonstrate that this type of analysis allows
us to obtain consistent results from isothermal and nonPhys. Chem. Chem. Phys., 2007, 9, 53595366 | 5359

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

View Article Online

Scheme 1

isothermal data, and also to get important correlations with


the rheological behavior during curing.

Experimental
Materials
Furfuryl alcohol (FA) is a light-yellow liquid (Mw = 98.10 g
mol1, b.p. = 170 1C, purity 498%). It was generously
supplied by TransFurans Chemicals BVBA (Belgium) and
was used as received. An anhydride, maleic anhydride (MA)
(Mw = 98.06 g mol1, m.p. = 5156 1C, purity 499%), was
used as acid catalyst to induce homopolymerization of FA and
was obtained from Aldrich Chemical Co. MA was totally
dissolved in water (50% aqueous solution) and was mixed in
FA solution at a weight ratio of FA : MA = 100 : 2 under
vigorous mechanical stirring. The fresh mixtures were immediately analyzed to avoid prepolymerization at ambient
temperature.
Experimental techniques
DSC measurements were performed on a Mettler-Toledo DSC
821e equipped with STARr software. Computations for
kinetic evaluation were performed using both internally-written software, regularly upgraded,1820 and the advanced
model-free kinetics STARr software option. Temperature
and enthalpy calibrations were done by using indium and zinc
standards. Volatilization of polycondensation by-products
occurs during the curing of FA and, consequently, highpressure stainless steel crucibles (ME-51140404) were used
instead of aluminium crucibles because they can withstand a
vapour pressure of up to 15 MPa. Samples of approximately
10 mg were placed into sealed pans. The DSC measurements
of FA polymerization were conducted at the heating rates of 1,
2, 4 and 6 1C min1. A blank run was subtracted for each
heating rate experiment. The DSC runs under isothermal
conditions were carried out at 110, 120, 130, 140, and
5360 | Phys. Chem. Chem. Phys., 2007, 9, 53595366

150 1C, respectively. The second heating of the cured samples


did not show any residual heat release.
Infrared spectroscopy (IR) was used to monitor the structural changes after dierent stages of polymerization. The IR
spectra were recorded on a Perkin Elmer Spectrum BX II
spectrophotometer. The attenuated total reectance (ATR)
mode was used to characterize starting monomer (FA), PFA
prepolymer in the gelled state and also the PFA thermoset
(totally cured polymer). To prepare samples for IR, a rst
DSC scan was performed to determine the relationship between a and T. Then, the prepolymer in the gelled state was
obtained by directly curing the FA monomer in DSC crucibles
at 1 1C min1 to a temperature of 113 1C, corresponding to a
B 0.54. Then, the sample was quickly quenched to 10 1C.
The same procedure was applied to obtain the totally cured
polymer (heated to T = 220 1C, a B 1).
Rheological measurements were conducted on oscillating
mode with parallel plate geometries (40 mm diameter and
1 mm gap) of a Bohlin C-VOR rheomether with strain
convection heating. The linear viscoelastic range of the material at its liquid and solid state was evaluated by a strain sweep
to determine a deformation that can be used to measure
complex viscosity during the whole polymerization process.
Measurements were carried out on auto stress mode with a
frequency of 1 Hz, and a deformation of 0.05%. Time to
gelation was estimated as the point where storage (G 0 ) and loss
modulii (G00 ) curves intersect. The polycondensation reaction
is accompanied by a water release, leading to the formation of
empty spaces between the plates. Due to this modication of
geometry during experimentation, the measured viscosities are
relative values.
Theoretical calculations
The overall rate of reactions is commonly described by the
following equation:21
da
kTf a
dt
This journal is


c

the Owner Societies 2007

View Article Online

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

where da
dt is the reaction rate, a is the conversion degree, k(T) is
the rate constant, t is the time, T is the temperature, and f(a)
is the reaction model. The dependence of the rate constant
is described by the Arrhenius law:


E
kT A exp 
2
RT
where E is the activation energy, A is the preexponential factor
and R is the gas constant.
The reaction rate is generally determined by the rates of
both chemical reaction and diusion22 The ratio of the
characteristic times of chemical reaction and diusion determines which of these two processes is rate limiting. According
to Debye,23 the relaxation time for a molecule is directly
proportional to viscosity:
t

4pa3 Z
kB T

where EZ is the activation energy of the viscous ow, Z0 is the


preexponential factor. Eqn (4) usually holds for liquids well
above their Tg. At temperatures closer to Tg, the temperature
dependence of viscosity follows the WLF equation.25 The
diusion activation energy, Ed can be obtain as follows
d ln kd
EZ RT:
dT 1

If RT is much smaller than EZ, Ed is practically equal to EZ,


which was experimentally found by several workers.24 Then,
substitution of the Arrhenius equation for k gives the apparent
activation energy, Ea:


d ln kef
EZ RTk EkZ

6
Ea R
dT 1
k kZ
where E is the activation energy of a chemical reaction. Eqn
(6) suggests that depending on the temperature, the apparent
activation energy, Ea, may take values between E and EZ.
Furthermore, the Ea dependence should be dierent for the
cures performed under a dierent temperature range.
The simplest isoconversional method proposed by Friedman26 allows the activation energy E to be determined for each
given conversion degree, and can be expressed as:
 
da
Ea
ln
lnAa f a 
7
RTa;i
dt a;i
where the subscript i denotes the ordinal number of a nonisothermal experiment conduced at the heating (or cooling)
rate bi. Eqn (7) allows the activation energy E to be determined
for each given conversion degree (a). The heat ow measured
in DSC is proportional to both overall heat release and cure
rate. In dierential scanning calorimetry (DSC), the conversion degree is dened as the ratio between the heat exchanged
This journal is


c

the Owner Societies 2007

The quantity (dH/dt)i represents the heat ux measured in


DSC at time ti, t1 is the time corresponding to the rst
integration bound and t2 to the second integration.
An advanced isoconversional method has been developed in
order to be applicable for any temperature programming, and
was used in this study.27,28 According to this method, for a set
of n experiments carried out at dierent heating programs,
Ti(t), the activation energy is determined at any particular
value of a by nding the value of Ea that minimizes the
function
FEa

n X
n
X
JEa ; Ti ta 
i1 jai

where Z is the viscosity of the medium, a is the molecular


radius, kB is Boltzmanns constant. This suggests viscosity as a
crucial factor of diusion control. Eyrings theory of viscosity24 gives rise to the Arrhenius type of temperature dependence:
 
EZ
4
Z Z0 exp
RT

Ed R

at time ti (Hi) and the total heat released by the reaction (Q):
R ti
Hi
t dH=dti dt
R t12
:
8
ai
Q
t1 dH=dti dt

JEa ; Tj ta 

Henceforth, the subscript a denotes the values related to a


given extent of conversion. In eqn (9), the integral:
JEa ; Ti ta  

Zta



Ea
dt
exp
RTi t

10

ta Da

is evaluated numerically for a set of experimental heating


program. Integration is performed over small time segments,
allowing for the elimination of a systematic error that occurs
in the usual integral methods when Ea varies signicantly with
a. In eqn (10), a is varied from Da to 1 Da with a step Da =
m1, where m is the number of intervals chosen for analysis.
The integral, J in eqn (10) is evaluated numerically by using
the trapezoid rule. The minimization procedure is repeated for
each value of a to determine the Ea-dependence.29 An advantage of the advanced isoconversional method is that it employs
exactly the same computational algorithm to evaluate the Eadependence from both isothermal and nonisothermal data.
The apparent activation energy calculated by isoconversional
methods is a global energy that may include several chemical
reactions (multi-step kinetics) as well as physical transformations (evaporation, gelation, vitrication). These methods can
take into account for the change from a chemically controlled
reaction to a reaction controlled by the change in viscosity or
by diusion.20,34,35,37

Results and discussion


Evolution of structure during FA polymerization
Partial or complete polymerization of the FA monomer has
been performed in order to study the evolution of structure
during polymerization and to highlight the DA reactions. Fig.
1 shows the IR spectra of starting FA monomer, those of PFA
prepolymer at the gelled state and also the PFA when completely cured. The comparison of the three spectra reveals
some important band-modications as a consequence of polymerization. As observed in Fig. 1, the polymerization leads to
a strong decrease of the OH stretching band at 3450 cm1
and those corresponding to the furanic bands around 1500,
1015, 920 and 880 cm1. A new band develops at 1560 cm1
Phys. Chem. Chem. Phys., 2007, 9, 53595366 | 5361

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

View Article Online

Fig. 1 IR spectra of FA monomer (a), PFA prepolymer (b) and PFA


totally cured polymer (c).

and can be attributed to a CQC skeletal stretch vibration in


the furan ring interconnected to the furanCH2furan group
after formation of oligomers by a condensation reaction
(Scheme 1a). The formation of chromophores during polymerization (Scheme 1b) is accompanied by an increase in the
relative intensity of the QCHdihydroxyfuran rings stretching band at 3150 cm1 together with the decrease of the
furanCH2furan stretching band at 2930 cm1. The evidence
of DielsAlder cycloaddition is marked by the appearance of a
non-aromatic CQC weak band at 1670 cm1 in the PFA
thermoset. The occurrence of the carbonyl signal at 1712 cm1
points out the formation of a ketonic structure by a furan ringopening reaction. However, these are considered side reactions, and lead to intermediate products that do not inuence
the overall rate of polymerization.30
Non-isothermal measurements
Normalized non-isothermal DSC curves of FA polymerization
are shown in Fig. 2. The heat released during non-isothermal
polymerizations of FA was combined and traced with the
corresponding variation of the extent of conversion with
temperature (Fig. 2). The non-isothermal DSC curves were

Fig. 2 DSC data (solid line) of the heat released during nonisothermal polymerizations of FA and the corresponding variation of extent
of conversion with temperature (dashed line). The heating rate of each
experiment (in 1C min1) is indicated by each curve.

5362 | Phys. Chem. Chem. Phys., 2007, 9, 53595366

Fig. 3 Variation in the eective activation energy with conversion


obtained for FA polymerization under nonisothermal conditions b =
16 1C min1 (open triangles) and evolution of the complex viscosity
at 4 1C min1 (solid line).

used to estimate the total heat release of reaction, Q. Integration of the DSC peaks resulted in decreasing Q values of 709,
685, 620, and 593 J g1 for the respective heating rates 1, 2, 4,
and 6 1C min1.
Milkovic et al.17 observed two distinct exothermic DSC
peaks for the polymerization of FA with p-toluene sulfonic
acid. These stages were attributed to the extension of the
polymerization reaction, followed by a distinct stage of crosslinking. As seen in Fig. 2, a single exothermic DSC peak is
observed in the case of FA polymerization. These DSC data
were used for the evaluation of the dependence of the apparent
activation energy, Ea, on conversion, as shown in Fig. 3. In
parallel is plotted the variation of the complex viscosity in
order to highlight rheological changes during polymerization.
According to this Ea-dependence, it can be concluded that the
non-isothermal acid-catalyzed polymerization follows multistep kinetics expressed by dierent apparent activation energies. The obtained Ea values are in agreement with the data
previously reported by Milkovic et al.17 of 73 and 106 kJ
mol1 using the Kissinger method. The Kissinger method
allows the computation of a single activation energy value
based on the evolution of the temperature of the peak
maximum with the heating rate.
At the beginning of the reaction, (a o 0.10) the Ea values
decrease from 70 to 55 kJ mol1. Such a decrease of Ea during
the initial cure stages has already been observed for an epoxynovolak system in a previous work.31 This behaviour was
explained by a diusion controlled kinetic due to the high
viscosity of the medium and Ea was correlated to EZ. However,
the system under investigation had a higher initial viscosity
(B102 Pa s), and the viscosity decreased strongly with increasing temperature in the initial stages (a o 0.20). In other
cases,32 the Ea decrease during the initial cure stages was
interpreted by a control of an autocatalytic step. According
to Fig. 3, the viscosity of the medium has a low initial value,
corresponding to the viscosity of the small FA oligomers and
remains constant until a B 0.100.15, in perfect agreement
This journal is


c

the Owner Societies 2007

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

View Article Online

with the Ea decrease. This quasi-constant viscosity value


results from the competition between the temperature-dependence of viscosity and the increase of molecular weight. Thus,
this suggests that the decreasing shape of Ea cannot be
attributed to a diusion-controlled rate but rather to an
autocatalytic mechanism that corresponds with the formation
of the active species for the rst stage of condensation. Under
acidic conditions, the chain reaction starts by the formation of
a furyl carbenium active center able to react with the monomer.33 This active center is regenerated at every step, allowing
the reaction to continue. The viscosity increases when sucient active species are generated and condensation reactions
start to become predominant (a B 0.15; this marks the end of
control by an autocatalytic step).
For 0.10 o a o 0.40, Ea values are around 60 kJ mol1 and
quasi-constant on conversion, suggesting that a single reaction
dominates in this conversion range.34 According to previous
studies35,36 the value of 60 kJ mol1 is typical for condensation
reactions of hydroxymethyl groups into methylene phenolic
bridges in phenolformaldehyde (PF) systems. The accepted
mechanism of polycondensation of FA is close to that of PF.
Indeed, the polycondensation of FA starts by linear condensation of a hydroxymethyl group to a furan ring and formation
of a methylene furanic linkage (Scheme 1). According to this
mechanism, the Ea values can be attributed to self condensation reactions that occur predominantly until a B 0.40.
Formation of linear oligomers occurs for this stage which is
expressed by a quasi-exponential increase of viscosity of the
medium. This increase is generally observed for the curing of
thermosets before gelation. Additionally, water is an important by-product of this polycondensation process (B18.35%
w/w) and can have a transient plasticizing eect on viscosity
only in the early cure stage (0 o a o 0.30), where the FA
oligomers have still a relatively low molecular weight. Moreover, due to the high temperature of polymerization
(4100 1C), the water is rapidly vaporized from the mixture.
For 0.40 o a o 0.45, Ea increases suddenly from 60 to
about 70 kJ mol1, revealing a change in the reaction mechanism. In this range, the DielsAlder cycloadditions can start
between oligomers formed in the previous step. The kinetics of
DielsAlder cycloadditions have been extensively studied, and
the reported activation energies of furanic diene cycloadditions on dierent dienophiles are ranging about 68 kJ mol1
(UV-Vis),37 83.6 kJ mol1 (ab initio),38 101.2105.1 kJ mol1
(DSC),39 82.395.3 kJ mol1 (ab initio),40 77.3145.0 kJ mol1
(ab initio),41 72.286.6 kJ mol1 (RMN).45 These values are in
agreement with ours and, therefore, this change in the reaction
rate at a B 0.45 could be attributed to the [4 + 2] cycloadditions between the furanic diene and the dihydrofuranic dienophile. Endo and exo DielsAlder cycloadducts cannot be
distinguished because they are formed with very close activation energies42 (dierences of about B24 kJ mol1). Moreover, at this extent of conversion, the observed material
becomes dark-coloured due to conjugated sequences.6 Formation of DA cycloadducts is conrmed by the appearance of
specic peaks registered in IR spectra (Fig. 1). Development of
interconnections between cycloadducts is still correlated with a
continue increase of the viscosity medium in this region as
shown in Fig. 3.
This journal is


c

the Owner Societies 2007

Fig. 4 Variation of G 0 (open symbol), G00 (solid symbol) during non


isothermal cure at 4 1C min1 (circle) and isothermal cure at 120 1C
(triangles).

For 0.45 o a o 0.63, Ea decreases from 70 to 50 kJ mol1.


In this conversion interval, the gelation process takes place as
shown in Fig. 4. The gel point of the system, taken as crossover between G 0 and G00 as seen in Fig. 4, is about agel E 0.52.
At the gel point, the growing oligomer chains, obtained by
DielsAlder cycloadditions, reach an unique, giant macromolecule, by connections between the furanic rings. When gelation takes place, the molar mass becomes innite and the
viscosity increases. This implies a low molecular mobility
which induces the decrease of the overall reaction rate that
becomes controlled by diusion of unbranched small oligomer
chains. This is mostly due to an inherent increase in viscosity
after branching cycloadditions and results in a decrease of Ea.
A similar decrease of Ea after gelation has been previously
reported for the epoxy/amine crosslinking37 and has been
attributed to the beginning of a diusion-controlled rate. This
latter change in the kinetic regime is easily observable upon the
evolution of viscosity. Indeed, for 0.63 o a o 0.72, the
viscosity reaches a constant value (plateau). As explained
above for the initial cure stage, these apparent constant values
of viscosity demonstrate a complex temperature dependence,
which is a superposition of two phenomena. Firstly, at a
constant value of the molecular weight, viscosity decreases
with increasing temperature. Secondly, an increase in the
temperature accelerates the cure and promotes an increase in
molecular weight and viscosity.
For 0.63 o a o 0.85, the apparent activation energy
increases to about 100 kJ mol1 (a E 0.85). The variation of
the polymerization rate leads to a dramatic change of viscosity. We observe a nal exponential increase of viscosity that is
clearly indicating the continuation of crosslinking reactions.
Because the temperature increases linearly in non-isothermal
experiments, the chain mobility increases and the chemical
reactions can be reactivated.
Thus, the increase of Ea to about 100 kJ mol1 could be
explained by the contribution of further DielsAlder
Phys. Chem. Chem. Phys., 2007, 9, 53595366 | 5363

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

View Article Online

cycloadditions to the cure rate. In the gelled state, the previously-formed longer chains are branched together by further
cycloaddition reactions leading to three-dimensionally crosslinked materials. At this stage of the reaction, the rate of
crosslinking is limited by the mobility of longer polymer
chains and diusion encounters a large energy barrier due to
the cooperative nature of the motions, leading to higher Ea
values.20,43
For a Z 0.85, Ea reaches a maximum, 100 kJ mol1 and
starts to decrease at vitrication (a Z 0.90). The monomer
molecules become frozen in their positions in the glassy state
that results in a virtual cessation of the reaction. The curing
rate in the glassy state becomes controlled by diusion of
small, unreacted functions still present in the medium and
chemical reactions are considerably reduced.20 This means
that vitrication causes a dramatic decrease in molecular
mobility leading to a decrease of the eective activation energy
with increasing extent of reaction.
The DielsAlder cycloadditions have the particularity of
being thermally reversible3,6,1416 and retro DielsAlder reactions may occur at higher temperatures. However, according
to rheological measurements presented in Fig. 3, the viscosity
still continues to increase with the increasing extent of the
reaction and reaches a plateau at av E 0.90, indicating that the
system is in the glassy state. The vitrication of the polyfurfuryl alcohol network is conrmed by the observation of a
glass transition temperature obtained during a second heating
of the cured sample in DMA. As shown in Fig. 5, the
maximum of tan d (G00 /G 0 ) can be used as an estimate of glass
transition temperature at Tg(non-iso) E 115 1C. Such observations cancel the hypothesis of retro DielsAlder reactions,
because conversion back to oligomers (by scission of polymer
crosslinks) would have been accompanied by a decline of
viscosity as previously observed for certain furanic copolymers.16 Moreover, corroborating with the structural IR characterization, the formation of the chromophore is conrmed in
the nal thermoset PFA by continuous increase of the
QCHdihydroxyfuran rings stretching band at 3150 cm1
compared to PFA prepolymer (Fig. 1). The intensity of furanic

Fig. 5 Variation of tan d obtained by rheometric measurements at 4 1C


min1 of samples already cured under non-isothermal mode at 4 1C
min1 (solid line) and under isothermal mode at 120 1C (dashed line).

5364 | Phys. Chem. Chem. Phys., 2007, 9, 53595366

Fig. 6 DSC data (solid line) of the heat release during isothermal
polymerizations of FA and the corresponding variation of the extent
of conversion with time (dashed line). The temperature of each
experiment (in 1C) is indicated by each curve.

bands at 1500 or 1015 cm1 continues to decrease with


increasing extent of conversion conrming that no conversion
back to furanic oligomers has occurred through retro Diels
Alder reaction. Furthermore, the strong decrease in intensity
of the OH stretching band at 3450 cm1 reveals that the nal
product is totally cured.
Isothermal measurements
The isothermal DSC thermoanalytical curves are shown in
Fig. 6. A blank run has been carried out at each isothermal
temperature to evaluate the stabilization of the pan after
introduction into the furnace. The obtained isothermal DSC
data have been integrated to evaluate the total heat released of
isothermal reactions. The latter varies as 579, 560, 607, 620
and 647 J g1 for the temperatures of 110, 120, 130, 140, and
150 1C. For all isothermal measurements, a shoulder appears
in the rst part of the DSC curves indicating a change in the
reaction rate.
The advanced isoconversional method was applied on these
isothermal data and the resulting dependency associated with
viscosity changes were reported in Fig. 7. In isothermal mode,
the maximum heat released is about 647 J g1, while it is
709 J g1 (amax B 1) in the non-isothermal mode. In order to
compare isothermal and non-isothermal dependencies, absolute values of a were plotted in Fig. 7, instead of relative
values. This means that the value amax,iso B 0.91 is taken as the
maximum value reached in isothermal mode.
At the beginning of the reaction (a o 0.15), we observe a
slight increase from 65 to 75 kJ mol1 in Ea values that is
attributed to formation of the active species for the rst stage
of condensation. In comparison with non-isothermal experiments (Fig. 3), the Ea-dependence has an opposite behaviour
for this rst conversion interval. This can be explained by the
fact that in the non-isothermal mode the active center formation started at low temperatures and conversion. In the
isothermal mode, the active species appears, at a certain
temperature, in time. To compare isothermal and non-isothermal data, we must keep in mind that the temperature
domain ranges between 110150 1C for the former and
This journal is


c

the Owner Societies 2007

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

View Article Online

Fig. 7 Variation in the eective activation energy with conversion


obtained for FA polymerization under isothermal conditions T = 110,
120, 130, 140, 150 1C (open circles) and evolution of the complex
viscosity during cure at 120 1C (solid line).

between 60220 1C for the latter data. This could explain why
the autocatalytic stage is only predominant at the early stages
of the non-isothermal cure, corresponding to lower temperatures. In this conversion interval, the viscosity remains constant because only low molecular weight oligomers and water
are formed.
For 0.15 o a o 0.40, the general shape of the Ea-dependence is in agreement with non-isothermal data suggesting
that formation of linear oligomers by condensation reactions
is predominant for this stage. The apparent activation energy
of the polycondensation step appears to be around 75 kJ
mol1. In this region, the viscosity registered only a slight
increase. As suggested for non-isothermal conditions, water
has a plasticizing eect. The viscosity increases slightly because the growth of the chain length during condensation
reactions is compensated by the plasticizing eect of water
formed in the same reaction.
Due to the higher activation energy of DielsAlder cycloadditions, these rst branching reactions between linear chains
start at earlier conversion degree in non isothermal experiments. This explains why for this region, in non isothermal
experiments, we observe a more marked viscosity increase in
comparison with the isothermal mode.
As previously described for non-isothermal data, the
DielsAlder cycloadditions become predominant for 0.40 o
a o 0.45. This step is characterized by a sudden Ea-increase
associated with an exponential increase of complex viscosity.
In isothermal curing, the Ea value increases to about 87 kJ
mol1 for a B 0.45. The DielsAlder reactions become the rate
limiting step for a B 0.400.50. This conversion range corresponds to the region were the shoulders is observed on the
isothermal DSC curves (Fig. 6).
In isothermal mode the gel point is obtained for a B
0.500.52 as seen in Fig. 4 (T = 120 1C). In the range
0.45 o a o 0.70, Ea decreases to about 55 kJ mol1, suggesting that the cure kinetics become controlled by diusion of
molecules in the gelled media. At this stage, the molar mass
becomes innite and the viscosity increases exponentially.
This journal is


c

the Owner Societies 2007

Compared to non isothermal data, we do not observe a


plateau of viscosity in this region. In the isothermal mode
the viscosity is not temperature-dependent, and only the
increase of molecular weight is determinant on its evolution.
Then, as observed for non-isothermal data, Ea goes on to
higher values. This conrms our hypothesis that a second
cycloadditions branching stage may occur at this conversion
range (0.70 o a o 0.85) and is associated with an increase in
Ea. This second branching stage is also marked by an increase
of the slope of the viscosity that start for a B 0.70 which
corresponds to the minimum value of the Ea. Because in
isothermal mode the temperature remains constant, the chemical reactions cannot be reactivated by the temperature
increase and the completion of the cure is lower than in nonisothermal mode (amax,iso B 0.91). This explains why we do
not observe the Ea increase to such a high value of 100 kJ
mol1, as was seen in the non-isothermal mode. As shown in
Fig. 5 (second heating), a peak maximum is observed at 70 1C
and is attributed to the glass transition temperature, Tg(iso), of
the sample cured under isothermal condition. This is an
indication that the system has reached the glassy state after
isothermal curing at 120 1C. As expected, Tg(iso) o Tg(non-iso),
conrming a lower completion of cure in isothermal mode. As
already mentioned, this experimental results support our
hypothesis that the DielsAlder cycloadditions start at earlier
conversion degree in non isothermal conditions leading to a
higher degree of crosslinking.
For a 4 0.85, the system is in the glassy state and the Eavalues start to decrease, as described previously for nonisothermal data. Lower values of activation energy correspond
to diusion of small unreacted molecules in the glassy state. As
in non-isothermal experiments (Fig. 3), the viscosity reaches a
plateau indicating that vitrication occurs. Similar trends are
highlighted in both isothermal and non-isothermal mode.
The complementary nature of DSC and rheometric data
gives new insights into the overall reaction mechanism of these
complex polymerizations. A sudden increase of Ea reects a
change in the polymerization mechanism (condensations to
DielsAlder cycloadditions). Condensation reactions still continue to occur, even for a 4 0.40, but are not the rate-limiting
step of the overall reaction. After gelation, the viscosity plays
an important role in the polymerization kinetics. Usually, a
decrease of Ea corresponds to a diusion-controlled rate and is
associated with a plateau of viscosity (for gelation or vitrication phenomena), whereas an increase of Ea corresponds to a
chemical controlled rate and is associated with an increase of
viscosity.

Conclusions
The polymerization of FA upon the action of an acidic
catalyst is a complex chemical process that supposes a linear
condensation step and a branching reaction step leading to a
black thermosetting material. The curing kinetics of this
monomer have not been extensively studied, although the
applications of FA based resins are increasingly important.
Isoconversional analysis associated with rheological data lead
to the conclusion that the overall curing of FA follows multistep kinetics controlled alternatively by chemical reactions and
Phys. Chem. Chem. Phys., 2007, 9, 53595366 | 5365

Published on 08 August 2007. Downloaded by Georgia Institute of Technology on 05/11/2016 21:29:16.

View Article Online

diusion. Linear condensations are predominant at the early


curing stage and the DielsAlder reactions become rate-determining after the formation of the rst oligomers. These rst
cycloadditions cause an increase in the viscosity-inducing
gelation. At this stage of reaction, the overall cure rate
becomes controlled by the diusion of oligomers. Finally,
the latter stage is marked by an increase of apparent activation
energy, explained by a dierent stage of cycloadditions in the
gelled state, leading to the nal glassy materials. This stage of
reaction becomes controlled by diusion of longer polymer
chains, while diusion of small molecules in the gelled or
glassy state is associated with a decrease of Ea.
Nevertheless, this study also highlights some dierences in
the development of the mechanisms between isothermal and
non-isothermal polymerizations. The material cured in nonisothermal conditions presents higher degree of crosslinking,
leading to higher Tg values, because the DielsAlder cycloaddition starts at an earlier stage of conversion.

Acknowledgements
The authors would like to acknowledge Transfurans Chemicals, the partners of the ECOBINDERS project (http://
www.ecobinders.net) and the European Commission for
nancial support (ECOBINDERS FP6-2003-NMP-SME-3
project Furans and lignin based resins as eco-friendly and
durable solutions for wood preservation, panel, board and
design products).

References
1 A. P. Dunlop and F. N. Peters, The Furans, Reinhold, New York,
1953.
2 W. L. E. Magalhaes and R. R. de Silva, J. Appl. Polym. Sci., 2004,
91, 17631769.
3 A. Gandini and M. N. Belgacem, Prog. Polym. Sci., 1997, 22,
12031379.
4 G. Toriz, R. Arvidsson, M. Westin and P. Gatenholm, J. Appl.
Polym. Sci., 2003, 88, 337345.
5 L. T. Dao and E. Zavarin, Holzforschung, 1996, 50, 470476.
6 M. Choura, N. M. Belgacem and A. Gandini, Macromolecules,
1996, 29, 38393850.
7 M. Principe, P. Ortiz and R. Martinez, Polym. Int., 1999, 48,
637641.
8 R. Gonzalez, R. Martinez and P. Ortiz, Macromol. Chem. Phys.,
1992, 193, 19.
9 R. Gonzalez, J. Rieumont, J. M. Figueroa, J. Siller and H.
Gonzalez, Eur. Polym. J., 2002, 38, 281286.
10 R. Gonzalez, J. M. Figueroa and H. Gonzalez, Eur. Polym. J.,
2002, 38, 287297.
11 I. S. Chuang, G. E. Maciel and G. E. Myers, Macromolecules,
1984, 17, 10871090.

5366 | Phys. Chem. Chem. Phys., 2007, 9, 53595366

12 J. L. Philippou and E. Zavarin, Holzforschung, 1984, 38,


119126.
13 T. P. Schultz, Holzforschung, 1990, 44, 467468.
14 C. Moreau, M. N. Belgacem and A. Gandini, Top. Catal., 2004, 27,
1130.
15 R. Gheneim, C. Perez-Berumen and A. Gandini, Macromolecules,
2002, 35, 72467253.
16 C. Gousse, A. Gandini and P. Hodge, Macromolecules, 1998, 31,
314321.
17 J. Milkovic, G. E. Myers and A. Young, Cellul. Chem. Technol.,
1979, 13, 651672.
18 N. Sbirrazzuoli, D. Brunel and L. Elegant, J. Therm. Anal., 1992,
38, 15091524.
19 N. Sbirrazzuoli, Y. Girault and L. Elegant, Thermochim. Acta,
1997, 293, 2537.
20 N. Sbirrazzuoli, A. Mititelu-Mija, L. Vincent and C. Alzina,
Thermochim. Acta, 2006, 447, 167177.
21 R. B. Prime, Thermosets, in Thermal Characterization of Polymeric Materials, ed. E. A. Turi, Academic Press, New York, 2nd
edn, 1997, 1380.
22 D. A. Frank-Kamenetskii, Diusion and Heat Transfer in Chemical
Kinetics, Plenum Press, New York, London, 2nd edn, 1969.
23 P. Debye, Polar Molecules, The Chem. Catalog Co., New York,
1929.
24 T. Ree and H. Eyring, The Relaxation Theory of Transport
Phenomena in Rheology: Theory and Applications, Academic Press,
New York, 1958.
25 J. D. Ferry, Viscoelastic Properties of Polymers, John Wiley &
Sons, New York, 3rd edn, 1980.
26 H. Friedman, J. Polym. Sci., Part C: Polym. Symp., 1964, 6,
183195.
27 S. Vyazovkin, J. Comput. Chem., 1997, 18, 393402.
28 S. Vyazovkin, J. Comput. Chem., 2001, 22, 178183.
29 N. Sbirrazzuoli, L. Vincent and S. Vyazovkin, Chemom. Intell.
Lab. Syst., 2000, 54, 5360.
30 R. T. Conley and I. Metil, J. Appl. Polym. Sci., 1963, 7, 3752.
31 S. Vyazovkin and N. Sbirrazzuoli, Macromol. Chem. Phys., 2000,
201, 199203.
32 S. Vyazovkin and N. Sbirrazzuoli, Macromolecules, 1996, 29,
18671873.
33 A. L. Montero, L. A. Montero, R. Martinez and S. Spange, J. Mol.
Struct., 2006, 770, 99106.
34 N. Sbirrazzuoli, S. Vyazovkin, A. Mititelu, C. Sladic and L.
Vincent, Macromol. Chem. Phys., 2003, 204, 18151821.
35 G. He, B. Riedl and A. Ait-Kadi, J. Appl. Polym. Sci., 2003, 87,
433440.
36 G. He and B. Riedl, Wood Sci. Technol., 2004, 38, 6981.
37 D. J. Cott, K. J. Ziegler, V. P. Owens, J. D. Glennon, A. E.
Grahamb and J. D. Holmes, Green Chem., 2005, 7, 105110.
38 S. M. Bachrach, J. Org. Chem., 1995, 60, 43954398.
39 J. Bibiao, H. Jianjun, W. Wenyun, J. Luxia and C. Xinxian, Eur.
Polym. J., 2001, 37, 463470.
40 M. Avalos, R. Babiano, J. L. Bravo, P. Cintas, J. L. Jimenez, J. C.
Palacios and M. A. Silva, J. Org. Chem., 2000, 65, 66136619.
41 T. C. Dinadayalane, R. Vijaya, A. Smitha and G. Narahari Sastry,
J. Phys. Chem. A, 2002, 106, 16271633.
42 L. Rul s ek, P. Sebek, Z. Havlas, R. Hrabal, P. Capek and A.
Svatos , J. Org. Chem., 2005, 70, 62956302.
43 S. Vyazovkin and N. Sbirrazzuoli, Macromol. Rapid Commun.,
2006, 27, 15151532.

This journal is


c

the Owner Societies 2007

Vous aimerez peut-être aussi