You are on page 1of 10

Chemical Engineering Journal 173 (2011) 210219

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Chemical recycling of PET by catalyzed glycolysis: Kinetics of the


heterogeneous reaction
Mateus E. Viana, Andr Riul, Gizilene M. Carvalho , Adley F. Rubira, Edvani C. Muniz
Grupo de Materiais Polimricos e Compsitos, GMPC, Departamento de Qumica, Universidade Estadual de Maring, Av. Colombo 5790, 87020-900, Maring, Brazil

a r t i c l e

i n f o

Article history:
Received 29 March 2011
Received in revised form 5 July 2011
Accepted 20 July 2011
Keywords:
Glycolysis
PET
Granulometry
Surface area
Mathematical model
Glycolysis kinetic

a b s t r a c t
Polyethylene terephthalate post-consume (PET-pc) glycolysis was investigated by the use of ethylene
glycol (EG) and zinc acetate, as catalyst. It was focused the kinetic aspects through use of mathematical
model specially developed for application in this study. The grains-lot was sieved in different size ranges
and a relation between surface area and granulometry, surface area and temperature on the conversion
and depolymerization rate was proposed. At temperatures ranging 180190 C the depolymerization
rate is quite elevated and almost 100% of conversion is obtained up to 3 or 4 h reaction time. For lower
temperatures (170180 C), equilibrium is achieved and it becomes more important as the temperature
is decreased. The conversion prole showed an initial activation stage where the mass transfer between
the liquid and solid phases is minimal. The proposed mathematic model was based on these ndings and
on reaction mechanism that differentiates the reaction sites present in the PET surface. By that model the
value of rate constant (k) for each temperature, and the dependence of k with 1/T was calculated. Four our
best knowledge it is the rst time that a mathematical model considers the activation stage in the earlier
times of PET depolymerization reaction. The inputs yielding, time and temperature were included in the
used mathematical model that ts very well the experimental data obtained at temperatures higher than
180 C. This model helps to predict the necessary mass of PET for producing a desired amount of products.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Technological advances in the manufacture of PET allowed it to
be produced at low cost. Allied to its mechanical properties, this
has led to a signicant increase in the production of this polymer. According to the Brazilian PET industry association, ABIPET
(Associaco Brasileira da Indstria do PET) [1], the quantity and
proportion of PET that is recycled have been growing year-by-year
and currently 263 k Tons (55.6% of production) are recycled.
There are three main types of PET-recycling [26]: (a) chemical
recycling (or depolymerization); (b) quaternary recycling (energy
recovery); and (c) mechanical recycling. Chemical recycling such
as depolymerization by hydrolysis [79], alcoholysis and glycolysis [1013] have been recently gained much attention. For instance,
terephthalic acid (TPA) and ethylene glycol (EG) can be recovered
and used as raw materials in many industrial processes, including polymer synthesis while the bis-hydroxyethyl terephthalate
(BHET) can be used in synthesis of new PET or other co-polymers.
Lorenzetti et al. [2] focused on the resulting products in their
review of PET degradation methods. Their study shows that gly-

Corresponding author. Tel.: +55 44 3261 3664; fax: +55 44 3261 4125.
E-mail address: gizilenecarvalho@gmail.com (G.M. Carvalho).
1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.07.031

colysis has advantages over other degradation methods because of


the versatility of the resulting product (BHET). Important studies
have been published by Pardal and Tersac [1113] on glycolysis
of post-consume PET (PET-pc). In the rst study [11], the reactivities of different glycols used in the depolymerization of PET were
compared. In a later study [12], the authors evaluated the kinetics of heterogeneous glycolysis of PET with diethylene glycol (DEG)
at 220 C. The change in total mass in each phase was evaluated
and it was observed that initially there is an induction period (of
about 15 min) with minimum mass transfer between the phases.
Subsequently, the reaction accelerates (6090 min) and then the
reaction rate decreases. A gradual decrease in the PET molar mass
was demonstrated using size exclusion chromatography (SEC), and
the authors concluded that under the used conditions the monomer
to oligomer ratio remains nearly the same during the reaction but
at end of the reaction the monomers constitutes the larger fraction
of nal products.
Yoshioka et al. [14] studied the kinetics of the hydrolysis of
micronized PET catalyzed by nitric acid at temperatures from 70 C
to 100 C (heterogeneous reaction). The principal nding was that
the reaction rate depends on the effective area (proportional to
the fraction of unreacted PET). This model correlated well with the
experimental data, although, in this case, the operating conditions
of the reaction did not give rise to the induction period found in

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

Nomenclature
a
ABET
A0
Ap
Aps
Au
c
d
Dp
f(%)
K
k
k1
L
m(t)
m0
ma
mBHET
MBHET
MEG
moligomer
mprod
mtheo
mu
v
M
r
s
s
S0
S1
S2
St
ti
Vu
X
yi
[]

red

MarkHouwinkSakurada parameters
surface area from BET method
initial contact area between the PET-EG phases
surface area of the actual particle
area of a perfect sphere with volume equal to that
of the a actual particle
unitary area of a particle
bigger size of PET grains to 3 parameters model
proportionality constant
mean mesh size of the sieves
mass fraction of sample retained in a specic mesh
size of the sieves.
MarkHouwink parameters
rate constant of the reaction
rate constant of the reaction
mean thickness of PET grains
residual mass of solid PET as function of time
mass of PET added to the reactor and m is the
mean mass values for each sample
BHET mass
molar mass of bis(hydroxiethyl)terephthalate
molar mass of ethylene glycol
mass of oligomers
mass of products retained on the lter paper
theoretical mass of BHET calculated based on the
weight of reacts PET
mean mass of a grain
average viscosimetric molar weight
PET grains radius from geometrical two-parameter
model
standard deviation of the mass of the samples
standard deviation of the thickness of PET grains
carbonyl groups that do not generate reaction products when cleaved (the middle of the chain)
carbonyl groups that generate reaction products
when cleaved (end of the chain)
carbonyl group does not change the polymer
the total number of sites at the beginning of the
reaction (t = 0)
induction time
mean volume of a grain
percentage conversion of solid PET
ratio between the number of sites of a determined
species on the surface of the polymer
intrinsic viscosity
reaction yield
reduced viscosity
sphericity

the glycolysis by Pardal and Tersac [12]. In Yoshioka et al.s study,


the fastest reaction rates occurred at the beginning of the reaction,
when the contact surface between the phases remained greatest.
Wan et al. [8] proposed a kinetic model for the hydrolysis of PET
catalyzed by potassium hydroxide. In such model, the reaction rate
is proportional to the contact-area between the phases and to the
concentration of KOH (with the rst and second order hypotheses
tested). The rate of reaction is rst order with regard to both of
these factors. Higher reaction rate in the rst moments was found
by this model and was attributed to the greatest contact area and
also to the concentration of KOH.
Ruvolo and Curti [15] published the rst study relating the inuence of surface area on the alkaline hydrolysis of PET in ethylene

211

glycol solution. They compared the geometric area and the effective
surface area of PET, as measured by BET (BrunauerEmmettTeller)
analysis, and demonstrated that the effective area increases as the
reaction progresses. Based on this, the authors proposed a kinetic
model in which the effect provoked by the decrease in geometric
area, due to the reaction, on the reaction velocity is considered.
The glycolysis of PET has been object of study from different
point of view. Although the inuence of variables such as time,
temperature, EG:PET molar ratio, nature of catalyst, concentration,
particle size, stirring rate, reaction time on the glycolysis process
have been investigated [1517], available kinetic models do not
cover all aspects of depolymerization process. The method of catalyzed glycolytic depolymerization with optimization technique
was described by Goje and Mishra [16]. The authors pointed that
procedures and resulting kinetic parameters vary with assumed
kinetic model and applied data tting procedure. The different values of activation energy (Ea ) cited in literature for depolymerization
of PET was attributed to the changes in reaction parameters and to
different chemicals employed for PET depolymerization. According
to Paszun and Spychaj [6], the literature related to the glycolysis of
PET covers mainly the application of the resulting products, while
only few authors turn their attention to the reaction kinetics. When
glycolysis is carried out below the melting temperature of PET, the
reaction medium consists, initially, of a solid phase (pure PET) and
a liquid phase (EG + catalyst). However, as the reaction proceeds,
other phases appear: swollen PET, a solution of polyesters and oligoesters, until at the end of the reaction there is only a liquid phase
(solution of glycols and oligoesters) [12]. The depolymerization
changes from a heterogeneous reaction to a homogeneous reaction, as the reaction progresses. Lpez-Fonseca et al. [17] developed
a theoretical model to predict the time conversion of PET during
glycolytic depolymerization. The authors observed that at initial
stages the reaction occurred in a heterogeneous phase and only at
higher reaction times the reaction became a single homogeneous
phase. The kinetic model was developed according to a homogeneous reversible catalytic model and was found to be consistent
with experimental data in the range temperature of 150196 C.
If the depolymerization reaction occurs initially on the surface of
the PET particles, what is the inuence of contact-area between the
phases (that is, the surface area of the PET grains)? The progress of
the reaction depends on the EG diffusing onto the surface of the
PET and on the removal-rate of the depolymerized material from
the surface of the PET (dry and swollen) into the solution. Hence,
the diffusion process can be taken to control the reaction rate and,
accordingly, the rate at which the solution is stirred also becomes
an important parameter in the reaction mechanism.
In one of their studies, Pardal and Tersac [12] examined the inuence of temperature, the presence of a catalyst and the morphology
of the PET. The observation that reactivity is much greater at 220 C
suggests that the diffusion of diethylene glycol in the PET is favored
at this temperature, increasing the reaction rate as compared to
reactions at lower temperatures.
With the objective of contributing to the understanding of the
glycolysis reaction mechanism for PET-pc with EG, this study proposes a new kinetic model for this reaction. The proposed reaction
mechanism is based on the possibility that the PET chain is cleaved
at different sites by the EG and that the amount of cleavages
depends on the ratio of the number of sites on the surface of the PET
grains. The following steps were performed to achieve this objective (i) determination of the geometric surface area of the PET-pc
grains using geometric models with two and three parameters; (ii)
experimental determination of conversion (X) as a function of the
granulometry of the grains of PET-pc; (iii) study of the inuence
of time and temperature on the heterogeneous depolymerization
reaction; (iv) determination of the PET-pc conversion (X) using the
proposed model, for different time and temperature conditions, and

212

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

comparison with the experimental results; (v) characterization of


the products obtained in the PET depolymerization reaction by DSC.
2. Experimental method
2.1. Materials
EG and zinc acetate (catalyst) were acquired from Synth (Brazil)
and used without purication. PET-pc, from soft drinks bottles, supplied by the company Plaspet Reciclagens Ltda (Maring, Brazil), was
washed and dried in an oven at 50 C to constant mass.
Viscosity measurements were made at 25 C in a 1:1 solution of 1,2-dichlorobenzene/phenol (m/m) to estimate the
viscosity-average molecular weight of PET. The equation due
MarkHouwinkSakurada
va
[] = K M

(1)

was used, being the values of K and a equal to 0.469 103 dL g1


and 0.68, respectively [18]. The intrinsic viscosity value found was
v ) cal[] = 0.78 dL g1 . The viscosity-average molecular weight (M
culated for the polymer was 54.600 g mol1 .
2.2. Analyses
To study the inuence of the size of the PET particles on the
kinetics of the reaction, a granulometry test was carried out using
a series of Tyler standardized sieves [19]. To do this, samples of
PET, washed and dried in an oven at 50 C to constant weight, were
vibrated for 20 min in a sieves set to split them into six size ranges.
The weight distribution as a function of mean particle size was evaluated by weighing the fraction retained by each sieve. The mean
mass of a grain (mu ) was estimated by weighing 15 random samples containing n = 30 grains each. The value of mu was calculated
from the mean mass, as given in equation

30

mu =

n (mu )i
i=1 i
30
n
i=1 i

30

i=1
30
n
i=1 i

(2)

The mean volume of a grain (Vu ) was estimated based on the


nominal density of PET (1.375 g cm3 ) and its mean mass, mu .
The Grubs test was applied to the data to reject anomalous data
[20], after which the values of mu and Vu were estimated for each
size range. As the mean thickness (L) of particulate PET is limited
by the thickness of the bottles, a digital micrometer was used for
measuring this axis for each granulometry. Thus, two hypothetical models related to the characteristic geometric-thickness of this
material could be proposed and evaluated.
The surface area was measured by means of BET adsorption
isotherms using a Quantachrome instrument, model NOVA-1000.
The mean grain area values (ABET ) as a function of granulometry
were used to evaluate the best geometric model further utilized
for calculating the area of a single PET particle.
The glycolysis reactions were designed to produce conversion
curves as a function of a given variable keeping the others constant. The variables studied were: time, temperature and particle
size. The mass of PET-pc used for each run was 15 g (78 mmol repeat
units) and the initial amount of the zinc acetate catalyst was 1.76 g
(8 mmol), giving a mol ratio of 9.75:1 for PET:zinc acetate. The
amount of EG, 60 g, was used thus in great excess to avoid problems
of homogenization in the reaction batch. The reactions were carried
out at atmospheric pressure using a three-necked ask equipped
with magnetic stirring, heating, a thermometer and a reux condenser. EG and the catalyst were weighed and added to the reactor,
and then heated to a requested temperature. Simultaneously, the
PET-pc was heated to the same temperature, in an oven, and then
quick transferred to the reactor. A stopwatch was triggered at the

moment the PET-pc was added. When a specied reaction time


had been elapsed, the heating was removed and boiling water was
slowly added to the system. Next, the contents of the reactor were
ltered (rst ltration) using a Tyler series sieve with a mesh size of
80 to collect the unreacted PET (PET-NR), and more boiling water
was used to remove the product eventually adhered to the PET.
The total volume of water added was 300 mL for each run. The ltrate was cooled to 4 C for precipitating the glycolysis products,
which were further ltered (second ltration) using quantitative
lter paper, and dried in an oven to constant weight. The ratio
between the mass resulting of second ltration products (mprod )
and the theoretical mass of BHET obtained (mtheo ), in accordance
with the stoichiometry of the reaction, was used to analyze the
reaction yielding by the equation
=

mprod
mtheo

(3)

The conversion (X) of solid PET was calculated using


X=

m0 m(t)
m0

(4)

where m0 is the mass of PET added to the reactor and m(t) is the
remaining solid PET mass as a function of time.
The relationship between conversion and the PET-pc geometric
area was evaluated for the glycolysis reaction at 180 C for 90 min.
A plot of conversion against time at the different temperatures for
A0 = 593.5 cm2 was produced and analyzed.
Thermal characterization of the samples was performed using a
Shimadzu DSC 50 calorimeter, with a heating rate of 10 C min1 in
an atmosphere of nitrogen at a ow rate of 20 mL min1 . The DSC
curve obtained for the product from second ltration was compared with the DSC curve for the oligomeric diols derived from
terephthalic acid [18].
3. Mathematical modeling
3.1. The reaction mechanism and denition of variables
In this study, a kinetic model with equations based on the different possibilities for polymer chain cleavage by EG is proposed.
The following hypotheses were considered for building the model:
(a) The reaction between an EG molecule and an ester group located
at the surface of the polymer causes a cleavage of polymer chain
at the point of the reaction. The reaction occurs at the interface
between the solid PET-pc and the diffused EG.
(b) If the reaction occurs between an EG molecule and an ester
group situated close to the end of PET chain, the each cleavage contributes to the reaction progressing and a leaving group
is formed consisting of a monomer (BHET) or an oligomer of low
molar mass.
(c) If the reaction occurs between an EG molecule and an ester
group situated far from the end of the chain, the cleavage does
not contribute to the reaction progress. In this case, the polymer molecule splits into two, forming two new chain-ends at
the point of cleavage. If this happens, the mass of the polymer is
increased by the accommodation of a molecule of EG relative
to the mass of polymer chain before the cleavage.
The sites where reaction may take place, that are the ester
groups located at the surface of the polymer, are classied as: S0 ,
S1 and S2 . The S0 sites indicate carbonyl groups that do not generate reaction products when cleaved (situated at the middle of the
chain). The S1 sites indicate carbonyl groups that generate reaction
products when cleaved (situated at the end of the chain). The S2
sites are inert, as a reaction involving these carbonyl groups do not

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

change the mass of polymer chain. In this case the intermediate


resonance structure would contain two identical leaving groups,
which are two ethylene glycol molecules.
To simplify the model, only BHET is considered to be a leaving
group. Hence, there are two S1 -type sites next to each end of the
chain. These considerations are illustrated in Fig. 1.
Another important denition to be considered in the proposed
mathematical model is the ratio between the number of sites of a
determined species on the surface at given reaction time, t, and the
number of sites at the beginning of the reaction, given by
S (t)
yi = i
St

(5)

where yi is the ratio of sites of type i. Si (t) is the number of sites


of type i located on the surface of the polymer, and St is the total
number of sites at the beginning of the reaction (t = 0).
At t = 0, almost all of the sites on the surface are S0 -typed, such
that the rst cleavages do not contribute to the conversion of the
polymer, but to the creation of chain ends at the polymer surface
(generating S1 sites). Accordingly, there is an increase in the yS1
ratio and a decrease in the yS0 ratio, as shown in the schema of
Fig. 2.
As a consequence of the yS1 growing ratio, an increase in the
number of effective cleavages occurs, that is, the cleavages that
generate leaving groups rise (reaction products).
Accordingly, the solid phase mass increases with each cleavage
at S0 due to the absorption of EG, and decreases with cleavages at
S1 due to the formation of BHET because EG is also absorbed in this
situation.
3.1.1. Denition the differential equations
The material balance equation (instantaneous mass balance) to
a dynamic system gives the following equation [9], being the units
of each term in Eq. (6) [mass time1 ].
dm
in m
out
=m
dt

(6)

The terms for the rates of mass entering and leaving will be substituted by differential or algebraic equations, as required for each
specic case.
Mathematically, the proposed mechanism can be described by
the material balance, as given in Eq (6), where the solid PET granule
is considered as the dynamic system. According to this equation,
the change in the PET granule mass over time (dm/dt) equals the
difference between the mass entering in the granule and the mass
in ) rises as EG
leaving out the granule. The rate of mass entering (m
attacks the S0 sites. Taking this rate as being of rst order in relation
to yS0 and to mass of solid residual PET, gives the equation
in = kyS0 (t)m(t)
m

(7)

where k is a rate constant dependent on temperature and represents the cleavage rate.
out ) as EG attacks the S1 sites,
The rate of mass leaving rises (m
at which time the EG molecules condense onto the polymer and
simultaneously soluble BHET molecules leave out of PET chain. Taking the leaving rate to have a rst-order relationship with yS1 and
to mass of solid residual PET, gives the equation
out = k1 yS1 (t)m(t)
m

(8)

where k1 is the rate constant of the reaction, also temperaturedependent.


Taking the cleavage velocity to be independent of the site type,
the probability of cleavage is greater for sites existing in greater
numbers. A cleavage at S0 site produces an increase on polymer
mass equal to the mass of EG molecule, and a cleavage at S1 site
produces condensation of one EG molecule into the polymer while,

213

simultaneously, the one BHET molecule leaves. Accordingly, the


ratio between the rate constants k and k1 is given by
MEG
k1
= 0.32299 = d
=
MBHET MEG
k

(9)

Substituting Eqs. (7)(9) into (6), gives


dm(t)
dt

= [kyS0 (t) dkyS1 (t)]m(t)

(10)

To nd a solution for this equation, it is necessary to express


the variables as a function of time, in order to generate an analytical solution for the conversion as a function of time. This makes
possible to evaluate the effect of temperature on the rate constants.
Expressing m(t) as a function of X.
The following equation
m(t) = m0 (1 X)

(11)

arises immediately from Eq. (4). Differentiating this equation and


substituting into Eq. (10), gives
dX(t)
= [kyS0 (t) dkyS1 (t)][1 X(t)] at X(0) = 0
dt

(12)

Expressing yS0 and yS1 as functions of time.


As the term is proportional to the S0 cleavage rate, the following
hypothesis can be written:
dyS0 (t)
= kyS0 (t) at yS0 (0) = 1
dt

(13)

In accordance with the illustration given in Fig. 1, taking into


account that each attack on an S0 site will form four S1 sites, the
following differential equation is proposed
dyS1 (t)
= 4kyS0 (t) at yS1 (0) = 0
dt

(14)

Hence, three ordinary differential equations (ODE) with their


respective contour conditions have been dened and were used
to obtain an analytical solution for the conversion as a function of
time.
4. Results and discussion
4.1.1. Granulometry and the surface area of PET-pc grains
The granulometric characterization of the PET-pc under investigation gave seven different particle sizes (Table 1). The term Dp is
the mean mesh size of the indicated sieves. The sample of the fraction at the bottom of the sieve set (7th sieve) was excluded from
the other analyses due to the low volume of this sample. The particles in all size ranges presented very different axis sizes. As the
thickness of the grains is limited to the thickness of the PET bottles,
so it was not possible to take these as spherical particles with an
average diameter equal to Dp , which is a common simplication
used for characterizing solid particles [19].
As it is useful to relate the area to the mean size of the particles,
two geometric models were tested. The rst geometric model (twoparameter model) assumes that the solid can be represented by
cylindrical particles, being the area dependent on the mean diameter and thickness of the particles, as shown in left side of Fig. 3. The
mean radius is calculated from the values obtained for the thickness and the volume of each particle. The second geometric model
(three-parameter model) assumes that the solid can be represented
by particles with a rectangular prole, as shown in right side of
Fig. 3. The shorter axis, c, is equal to the value estimated for L; the
intermediate axis, b, was taken to be equal to the value of Dp for
the particles; the longer axis, a, was calculated from the values for
Vu , a and b.

214

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

Fig. 1. Schematic representation of sites S0 , S1 and S2 in the PET molecule.

Fig. 2. Illustration of how S0 decreases and S1 increases during the conversion (X) of PET-pc.
Table 1
Estimates for mu and Vu and L as a functions of the granulometry from 15 samples of 30 grains each.
Dp (mm)

f (%)

ma (g)

s (g)

mu (mg)

Vu (mm3 )

L (mm)

s (mm)

7.18
5.56
3.56
2.03
1.44
0.89

5.43
25.15
60.62
6.16
1.80
0.77

1.1893
0.8212
0.6371
0.2140
0.1575
0.0921

0.1400
0.0243
0.0201
0.0176
0.0152
0.0117

39.64
27.37
12.74
3.567
2.230
1.149

28.83
19.91
9.267
2.594
1.622
0.835

0.5465
0.5014
0.4400
0.4027
0.3735
0.3401

0.0109
0.0093
0.0365
0.0200
0.0548
0.0346

Sample mass of 30 grains = ma ; Standard deviation of mass = s; Standard deviation of thickness (L) = s ; Mean mesh size of the shieves = Dp ; Unitary mass of grains = mu ; Unitary
volume of grains = Vu ; Percent of PET mass retained in the Tyler sieves f(%).

Another important characteristic of a given sample is its form


factor. The most common form factor for solid particles is the
sphericity ( ), a form factor based on the surface and the volume of particles [21,22]. The sphericity is dened as being the ratio
between the surface area of a perfect sphere with volume equal to
that of the particle and the surface area of the actual particle, given
by
=

Aps
Ap

(15)

where Aps is the area of the equivalent sphere; Ap is the area of


the particle. Accordingly, 0 < < 1, and = 1 for an ideal spheric
particles. Sphericity ( ) was calculated for the PET-pc grains from
the surface area data obtained using two geometric assumptions.
The two geometric assumptions take into account a thickness
limit for the particles. Table 1 gives the mean mass values for each
sample, ma , and the values calculated for mu , Vu and the standard
deviation (s). The increase in the PET thickness as a function of Dp

Fig. 3. Hypothetical models surface area. (a) Two-parameter model and (b) threeparameter model.

was attributed to the fact that it is more difcult to grind thicker


bottles. The capacity of the two hypothetical geometric models to
correlate the geometric area to the PET granulometry was tested
using these data.
4.2. Evaluation of the models used to calculate the surface area of
the PET grains
The estimated parameter values (Table 1) and the two hypothetical models (Table 2) were used to calculate the surface area
and the sphericity of the PET grains. For an additional analysis, the
surface area was measured using the BET method. However, instrument limitations allowed measurements of only the three smallest
size ranges. The instrument measures the specic area (Aesp = area
mass1 ) of the PET. Hence, the values were multiplied by mu to
calculate the unitary area (ABET ).
The values calculated for Au by applying the two models correspond to a geometric area, without taking surface roughness into
account. The values obtained for the sphericity are very low, providing evidence of the great surface irregularity of PET. For the area
determined using the BET method (ABET ), which does take surface
roughness into consideration, the values are greater than the calculated geometric areas; however, they must be proportional to
Au .
Aiming do a selection of best model, the Zingg Classication [23]
was used. It deals with the morphology of gross particles, separating
them into four classes (Fig. 4).
The values of p and q were calculated from the values of a, b and
c in the three-parameter model. For the lamellar PET grains (p and
q 2/3), the values of Au were estimated using the three-parameter

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

215

Table 2
Unitary area and sphericity as a function of granulometry.
Dp (mm)

7.18
5.56
3.56
2.03
1.44
0.89

2 parameters model

3 parameters model

r (mm)

Au (mm2 )

4.223
3.555
2.589
1.432
1.176
0.884

125.7
90.62
49.28
16.51
11.44
6.804

0.362
0.392
0.433
0.553
0.583
0.631

c (mm)

Au (mm2 )

7.804
7.142
5.916
3.173
3.015
2.761

127.5
92.26
50.46
17.07
12.01
7.397

ABET (mm2 )
0.357
0.385
0.423
0.535
0.556
0.580

52.00
36.57
20.48

Fig. 4. Zingg classication of different geometric particles.

A0 = nAu

(16)

where n is the number of particles present in the reaction medium,


estimated for each size range by the ratio between the total mass
(ma ) added to the reactor and the unitary mass (mu ). The results
are given in Table 3.
As expected, Au was greater for the larger size ranges, whereas
A0 was smaller for the larger size ranges.

90

Conversion (X, %)

model, while for the discoidal PET grains (p 2/3 and 2/3 < q 1.0),
the values of Au used were calculated using the two-parameter
model. To calculate the initial contact area between the PET-EG
phases (A0 ) for each size range, the values of the unit mass and the
total mass added to the reactor were used, as given in equation

60

30

4.3. Glycolysis reactions

400
The relationship between the conversion and the geometric
area of the PET-pc grains in the glycolysis reaction, (T = 180 C and
t = 90 min) were evaluated as a function of the initial contact area,
A0 , and are given in Fig. 5. As the reaction time was xed at 90 min,
it can be seen that the reaction velocity is largely dependent on the
contact area. As pointed by some authors [12,1517,24], stirring
rate inuence the conversion of PET. In conditions of high solution
stirring rate the mass transfer resistance can be eliminated in the
reaction medium. In all experiments the magnetic stirring is maintained constant at ca. 200 rpm and the mass transfer mechanism
was not eliminated neither considered in the mathematical model.
To estimate the PET mass required to produce a specic mass of
product (mprod ), the product yield was assessed. In accordance with
the stoichiometry of the reaction, 1.323 g of BHET (mtheo ) is formed
for every gram of PET reacted. However, due to inherent process
losses and to incomplete depolymerization, the mass of products
obtained on the lter paper, after drying, was smaller than the theoretical BHET mass, mtheo . The conversion (X) for the reactions was

600

800

1000

Initial contatct surface (Ao, cm2)


Fig. 5. Conversion of PET as a function of initial contact surface (reaction 90 min, at
180 C).

dened in Eq. (4). From this equation and take into account the stoichiometry of PET glycolysis to form BHET, the following equation
was obtained:
=

mprod
mtheo

mprod
1.323(m0 X)

(17)

The mass of products for each conversion was determined. All


of the values were normalized for an initial PET mass of 15 g. Subsequently, to calculate the mean reaction yield (), the mass of
products formed was compared with the theoretical value (mtheo ).
The results are given in Fig. 6.

216

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

Table 3
Values of p and q and the Zingg classication of the particles studied and the surface area for a 15 g sample of PET.
Dp (mm)

Class

Au (mm2 )

mu (mg)

A0 (cm2 )

7.18
5.56
3.56
2.03
1.44
0.89

0.072
0.090
0.124
0.198
0.259
0.382

0.920
0.779
0.602
0.640
0.478
0.322

discoidal
discoidal
lamelar
lamelar
lamelar
lamelar

125.7
90.62
50.46
17.07
12.01
7.397

39.64
27.37
12.74
3.567
2.230
1.149

376
553
1176
4258
6726
13055

473.3
501.6
593.5
727.0
808.0
965.7

25
100

80

75

%)

15

10

40
mprod

mtheo

0
0

20

40

60

80

20

X (%)

60

yield (

mass (g)

20

100

170 C
o
175 C
o
180 C
o
185 C
o
190 C

50

25

0
100

Conversion (X, %)

Fig. 6. Reaction yield in comparison with maximum theoretical yield (reaction


90 min, at 180 C).

50

100

150

200

250

Reaction time (min)


Fig. 7. Prole of PET-pc conversion with A0 593.5 m2 against time at different temperatures.

The left vertical axis on Fig. 6 gives the values of mprod and mtheo ;
the right vertical axis gives values of . The Grubbs test suggests that
the yield from the smallest conversion reaction should be excluded
when calculating the mean yield (as the conversion is small, any
loss results in large errors). Excluding the smallest conversion, the
mean reaction yield was 91.6%. Therefore, the mass of products
collected on the lter paper can be estimated, in terms of mean
value, using

area between the phases. Note that, for the used operating conditions, the induction period is greater at lower temperatures. This
behavior was not predicted by the proposed model and suggests
that diffusion may play an important role in the reaction mechanism allowing the reaction to take place more readily in the solid
phase at temperatures above 180 C.

mprod = 1.212m0 X

4.5. Mathematical modeling

(18)

4.4. Heterogeneous depolymerization kinetics for the glycolysis of


PET
The inuences of the reaction temperature and time on the reaction conversion were evaluated simultaneously for PET samples of
initial surface area of 593.5 mm2 , due the high volume of this sample. The graph of conversion as a function of time was plotted for
170, 175, 180, 185 and 190 C, as shown in Fig. 7.
The conversion rate (dX/dt) shows a delay at initial reaction
times (activation time), manly at 170 and 175 C. This behavior differs from that described in the literature for mathematical
models of PET degradation, but it agrees with the results of Pardal
and Tersac [12]. In the present study, this behavior was attributed
to the types of polymer surface sites. During the initial period of
the reaction, nearly all of the surface sites were of S0 type and the
rst cleavages did not contribute to the conversion of the polymer. During the induction time, cleavage of mainly this type occurs
and produces chain ends on the polymer surface (production of
S1 ). After a certain number of groups at S1 sites have been generated, the number of effective cleavages (at the S1 sites) increases,
increasing the conversion of PET-pc into reaction products (BHET),
followed by a subsequent decrease as the reaction progress. The
decrease in (dX/dt)T for longer reaction times can be explained by
the decrease in the solid phase and, consequently, in the contact

4.5.1. Analytical solution


The analytical solutions for the ODEs were obtained using the
computational tool Maple11TM . The rate constant was estimated
using the PolymathTM software. The differential equations and the
experimental data were inserted into the algorithm to obtain the
best t. To achieve correlation between k and T, a graph of ln(k)
versus 1/T was constructed for these 5 points, thus the parameters
for the Arrhenius equation were extracted from the plots of Fig. 8.
Note that a better t is obtained when two temperature ranges,
170180 C; and 180190 C, are considered. Considering only the
three higher temperatures (180, 185 and 190 C) the equation
k = 572.5 exp(5013.2/T )

(19)

is obtained. From this equation the calculated value for Ea was


41.7 kJ/mol. For the lower temperatures (170, 175 and 180 C), the
best t in Fig. 8 gives the equation
mprod = 1.212m0 X

(20)

is obtained and from this equation the Ea value calculated was


99.6 kJ/mol. The Ea values reported for PET depolymerization by
various researchers are different [1517]. The different values of
Ea were attributed to the change in reaction parameters and different chemicals employed for depolymerization of PET [16]. The
value of Ea calculated for lower temperatures, 99.6 kJ/mol, agrees

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

-4,4

y=-11977x+21.69
R2=0.990

-4,6

k=2.637.10 e-11977/T

-4,7
-4,8

lnk

-4,9
-5,0
-5,1
-5,2

-5,4

y=-5013.2x+6.346
2
R =0.994
-5013.2/T
k=572.5e
0,00218

80

40

0
170

-5,5
0,00216

induction time, t i (min)

120

-4,5

-5,3

217

0,00220

0,00222

0,00224

0,00226

1/T (K -1 )

180

190
o

temperature C
Fig. 10. Induction times (ti ) at different temperatures. The ti value was obtained, in
each case, from an intercept to the straight line in the major decomposition stage.

Fig. 8. Values of k as a function of 1/T. Fitting for lower temperatures (170 175 and
180 C, dash line) and higher temperatures (at 180, 185 and 190 C, solid line).

with the values reported in the literature, ranging between 85 and


100 kJ/mol, for the catalytic glycolysis of PET. The value of Ea calculated for higher temperatures, 41.7 kJ/mol, is less than values from
other researches. Therefore, this small value can indicate that the
glycolysis was rate determining by the kind of sites in the PET surface. In our case these two values of Ea obtained indicated that in
the conditions utilized in this work, the mechanism of reaction is,
probably changed as the temperature change from 170180 C to
180190 C. For the reaction to proceed EG must rst reach the surface of PET and then access to any site S1 . After the BHET molecule
be released from PET grain, new site is formed in the reminiscent
chain. Temperature affects the kinetic energy of molecules in solution, so that at low temperatures the low kinetic energy causes
the reaction rate depends on the mass transfer process. At higher
temperatures the average kinetic energy of molecules in solution
is sufcient to overcome this barrier and the reaction rate becomes
controlled only by the access of EG molecules to sites S1 . Thus
the model gives good results only under conditions where mass
transfer has no control over the process. The validity of the model
was conrmed by comparing the experimental data with the data
calculated for each of the temperatures. The values for k in the analytical solution were substituted for the relationships given above,
and the results given by the model are shown together with the
experimental data in Fig. 9(a) and (b).

These results point to the consistency of the model presented


here for 180 C and higher temperatures (Fig. 9(a)). This model
brings a new perspective on heterogeneous depolymerization, even
explaining the induction period rst reported by Pardal and Tersac [12]. However, the model was unsatisfactory at temperatures
below 180 C (Fig. 9(b)), as already mentioned.
As the glycolysis reaction occurs below the PET melting temperature, it is a heterogeneous reaction. In this case, the effect of the
EG diffusion mechanism at the surface of the PET-pc grains and the
diffusion of the reaction products in the solution must be considered in addition to the sites available for effective cleavage. Above
180 C, this is not the determinant effect and it is the number of
effective cleavages that controls the reaction velocity. The proposed
model ts the experimental data perfectly at these temperatures.
The proposed model does not t the experimental data for temperatures below 180 C. In this temperature range, the greater induction
period (Fig. 10) suggests that, apart from the numbers of available
sites for effective cleavage, the diffusion process must also be considered and that both determine the reaction rate. Then the Ea value
for reaction above 180 C is smaller than for reaction below 180 C.
The ti values decreased with increasing temperature, and
seemed to reach an asymptotic value. Since the system examined in
this study is a solidliquid heterogeneous reaction, the conditions
contacting a reaction, solvent to the solid polymer, signicantly
affect the reaction rate.

Fig. 9. Conversion of PET-pc as given by the experimental data (points) and the proposed model (line) as a function of temperature. (a) Temperatures of 180, 185 and 190 C
and (b) temperatures of 170 and 175 C.

218

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219

Heat flux rate /mW


Endotermic

BHET

Oligomers

PET-pc

100

200

Temperature /

300
o

Fig. 11. DSC thermogram for (a) PET-pc; products obtained in the second ltration:
(b) oligomeres retained on quantitative lter paper and (c) BHET obtained after
cooling the ltrated at 4 C.

4.6. Characterization of products


4.6.1. Thermal characterization
Depolymerization at intermediate operating conditions
(T = 180 C, t = 90 min, X = 55.9%) was conducted and the resulting
products were characterized. The solid mass obtained in the hot
ltration corresponded to 6.6% of the total product mass, and the
solid obtained in the cold ltration, corresponded to 93.4% of the
total product mass. The DSC curves for the hot and cold ltration
products are given in Fig. 11, with the PET-pc curve for comparison.
The peak at approximately 110 C in the DSC curve obtained for
the products of the rst (hot) ltration corresponds to the melting temperature of the BHET monomer. Less intense melting peaks
attributed to the presence of dimers and trimers can also be seen
at 151 C and 210 C. The peak in the region of 253 C is attributed
to the melting of the remaining solid PET, which then percolates
through the 180 m mesh into the products phase. The other peaks
were attributed to the vaporization of the materials and to impurities contained in the PET-pc that were retained in the rst ltration.
Although the values are not rigorously equal to those found for the
dimers in the literature, endothermic peak shifts were observed
when a mixture of products was analyzed because of the intense
interactions between the dimers [25].
For the material obtained in the second (cold) ltration, a rst
peak was observed at approximately 110 C corresponding to the
BHET melting range. The second peak was attributed to the vaporization of the sample above 255 C.
5. Conclusions
PET-pc is a granular solid with peculiar characteristics. The low
sphericity of the particles is an obstacle that restricts the particles
be treated as spheres. The approach used in characterizing the solid
demonstrated that the smaller particles can be treated as lamellar
particles and the larger particles as discoid particles.
Glycolysis was found to be efcient for PET-pc chemical recycling, as it can be conducted at atmospheric pressure and the
operating conditions required are relatively mild compared with
other methods. The reaction can achieve conversion percentages
close to 100% at temperatures above 180 C, when 78 mmol PET
repeat units is catalyzed by 8 mmol L1 zinc acetate. At lower
temperatures, there is apparent reaction equilibrium between
oligomers and unreacted PET, which leads to lower conversion.

An initial delay in the PET conversion can be seen at all temperatures. This induction period was attributed to the low initial
probability of the EG attacking the PET chain ends, which leads
principally to the formation of BHET.
Mathematical modeling and the reaction mechanism presented
here estimates the experimental data for temperatures of 180 C
and above. At these temperatures, the induction time observed can
be explained in terms of the types of sites (S0 ) on the polymer surface. For temperatures below 180 C, the diffusion process must
also be taken into consideration.
The product characterization showed that BHET is the main
reaction product, with small quantities of dimers and trimers also
being produced.
In the reaction yield study, the relationship established to calculate the mass of BHET formed as a function of the conversion,
showed that recovery was 91.6% of the theoretical value. This
percentage may be increased by implementing more efcient separation methods. However, the development of such methods was
not an objective of this study. Using this relationship and the solution from the mathematical model, it is possible to predict the
operating conditions and PET mass required to produce any quantity of products, thus providing a production planning tool for the
recycling of PET-pc and other polymers from a solidliquid reaction
such as the type studied here.

Acknowledgements
M.E.V. thanks to (CNPq, Brazil) for the master fellowship. The
authors thank CNPq, Brazil for the nancial support (proc. no.
309005/2009-4 and 481424/2010-5). All authors thank to COMCAP/UEM for access to DSC experiments.

References
[1] Associaco Brasileira da Indstria do PET (ABIPET)6 Censo da Reciclagem de
PET no Brasil, available http://www.abipet.org.br/, Accessed 01/16/2011.
[2] C. Lorenzetti, P. Manaresi, C. Berti, G. Barbiroli, Chemical recovery of useful
chemicals from polyester (PET) waste for resource conservation: a survey of
state of the art, Journal of Polymers and the Environment 14 (2006) 89101.
[3] V. Sinha, M.R. Patel, J.V. Patel, PET waste management by chemical recycling: a
review, Journal of Polymers and the Environment 18 (2008) 825.
[4] S.R. Shukla, A.M. Harad, L.S. Jawale, Recycling of waste PET into useful textile
auxiliaries, Waste Management 28 (2008) 5156.
[5] F. Awaja, D. Pavel, Review-recycling of PET, European Polymer Journal 41 (2005)
14531477.
[6] D. Paszun, T. Spychaj, Chemical recycling of poly(ethylene terephthalate),
Industrial & Engineering Chemistry Research 36 (1997) 13731383.
[7] G.M. Carvalho, E.C. Muniz, A.F. Rubira, Hydrolysis of post-consume
poly(ethylene terephthalate) with sulfuric acid and product characterization
by WAXD, NMR13C and DSC, Polymer Degradation and Stability 91 (6) (2006)
13261332.
[8] B.Z. Wan, C.Y. Kao, W.H. Cheng, Kinetics of depolymerization of poly(ethylene
terephthalate) in a potassium hydroxide solution, Industrial & Engineering
Chemistry Research 40 (2001) 509514.
[9] G. Gcl, T. Yalcnyuva, S. zgms, M. Orbay, Simultaneous glycolysis and
hydrolysis of polyethylene terephthalate and characterization of products by
differential scanning calorimetry, Polymer 44 (2003) 760976167.
[10] D. Navnath, Pingale glycolysis of postconsumer polyethylene terephthalate
waste, Journal of Applied Polymer Science 115 (1) (2010) 249254.
[11] F. Pardal, G. Tersac, Comparative reactivity of glycols in PET glycolysis, Polymer
Degradation and Stability 91 (2006) 25672578.
[12] F. Pardal, G. Tersac, Kinetics of poly(ethylene terephthalate) glycolysis by
diethylene glycol. I. Evolution of liquid and solid phases, Polymer Degradation
and Stability 91 (2006) 28402847.
[13] F. Pardal, G. Tersac, Kinetics of poly(ethylene terephthalate) glycolysis by
diethylene glycol. Part II: effect of temperature, catalyst and polymer morphology, Polymer Degradation and Stability 92 (2007) 611616.
[14] T. Yoshioka, T. Motoki, A. Okuwaki, Kinetics of hydrolysis of PET powder in nitric
acid by a modied shrinking-core model, Industrial & Engineering Chemistry
Research 37 (1998) 336340.
[15] A.C. Ruvolo-Filho, P.S. Curti, Chemical kinetic model and thermodynamic compensation effect of alkaline hydrolysis of waste poly(ethylene terephthalate)
in nonaqueous ethylene glycol solution, Industrial & Engineering Chemistry
Research 45 (2006) 79857996.

M.E. Viana et al. / Chemical Engineering Journal 173 (2011) 210219


[16] A.S. Goje, S. Mishra, Chemical kinetics, simulation, and thermodynamics of glycolytic depolymerisation of poly(ethylene terephthalate) waste with catalyst
optimization for recycling of value added monomeric products, Macromolar
Material Engineering 288 (2003) 326336.
[17] R. Lpez-Fonseca, I. Duque-Ingunza, B. de Rivas, L. Flores-Giraldo, J.I. GutirrezOrtiz, Kinetics of catalytic glycolysis of PET wastes with sodium carbonate,
Chemical Engineering Journal 168 (2011) 312320.
[18] J. Brandrup, E.H. Immergut, Polymer Handbook, third ed., John Wiley & Sons,
New York, 1967, VII-41.
[19] A.S. Foust, L.A. Wenzel, C.W. Clump, L. Maus, L.B. Andersen, Princpios de
Operaces Unitrias, second ed., Guanabara Dois, Rio de Janeiro, 1982.
[20] B. Barros Neto, I.S. Scarminio, R.E. Bruns, Como Fazer Experimentos: Pesquisa
e Desenvolvimento na cincia e na Indstria., second ed., Editora Unicamp,
Campinas, 2003.

219

[21] L.E. Sissom, D.R. Pitts, Fenmenos de Transporte, rst ed., Guanabara Dois, Rio
de Janeiro, 1996.
[22] A. Botari, L.D. Bernardo, Modelaco matemtica macroscpica da perda de carga
e da remoco de slidos suspensos totais na ltraco direta ascendente, Revista
Engenharia Sanitria e Ambiental 12 (2) (2007) 149159.
[23] T. Demur, R.P.D. WALSH, Shape and size characteristics of bedload transported
during winter storm events in the Cwm Treweryn stream, brecon beacons, 14,
Turkish Journal of Earth Sciences, South Wales, 2005, 105121.
[24] C.Y. Kao, B.Z. Wan, W.H. Cheng, Kinetics of hydrolytic depolymerization of melt
poly(ethylene terephthalate), Industrial & Engineering Chemistry Research 37
(1998) 12281234.
[25] S. Baliga, W.T. Wong, Depolymerization of poly(ethylene terephthalate) recycled from post-consumer soft-drink bottles, Journal of Polymer Science Part A
Polymer Chemistry 27 (6) (1989) 20712082.

You might also like