Sei sulla pagina 1di 9

Recovery of polyols from exible polyurethane foam by split-phase

glycolysis: Study on the inuence of reaction parameters


Carolina Molero, Antonio de Lucas, Juan F. Rodr guez
*
Department of Chemical Engineering, University of Castilla-La Mancha, Avda. Camilo Jose Cela s/n, 13004 Ciudad Real, Spain
Received 11 October 2007; accepted 30 November 2007
Available online 8 December 2007
Abstract
The general purpose of polyurethane chemical recycling is to recover constituent polyol, a valuable raw material. Among suitable processes,
glycolysis, especially in two phases, allows better quality products. Potassium octoate a compound derived from cycloaliphatic carboxylic acids
shows suitable catalytic activity.
A detailed study of the main reaction parameters affecting the reaction and properties of the recovered polyol has been carried out. They
include carboxylate catalyst concentration, reaction temperature and mass ratio of treated foam to the glycolysis agent.
An increase in the reaction temperature and catalyst concentration enhances the degradation rate, however, it also affects the process
negatively by promotion of secondary reactions and contamination of the polyol phase. Related to the glycolysis agent amount, the minimum
quantity required to split the phases has been determined, as well as the optimal ratio.
2007 Elsevier Ltd. All rights reserved.
Keywords: Polyol; Polyurethane; Recovery; Glycolysis
1. Introduction
Polymeric compounds with several hydroxyl groups,
known as polyols, are essential components in the manufac-
ture of polyurethanes (PU). Polyether polyols generally have
a molecular range of 180e8000 g mol
1
and are produced
by the polyaddition reaction of alkylene oxides like propylene
oxide (PO) and ethylene oxide (EO). By modifying molecular
weight and functionality, polyols provide a diverse range of
high performance properties in PU materials. Flexible PU
foams are the most important group among PU specialties,
reaching the 29% of the total production. They are widely
used in furniture, mattresses and automotive seats. Because
PUs are used in many everyday applications and industrial
uses, their wastes cause economical and environmental
problems. Such wastes comprise not only post-consumer
products but also scrap from slabstock manufacturing, which
can reach the 10% of the total foam production. An alternative
approach to landlling, is chemical recycling to convert the
PU back into its starting raw materials, especially polyols.
Hydrolysis, treatment with esters of phosphoric acid, ami-
nolysis with low weight alkanolamines and glycolysis have
been described as suitable procedures to break down the PU
chain [1e6] by transesterication. Most of these processes
produce a liquid mixture of products containing the recovered
polyol and other hydroxyl active groups, which can be used
only in blending with raw materials. Nevertheless, better
quality products can be achieved from exible PU foams using
a two-phase glycolysis, enabled by the higher molecular
weight of polyols used in this kind of PU. By means of an
excess of glycolysis agent, much larger than the stoichiometric
quantity, the reaction product splits in two phases, where the
upper layer is mainly formed by the recovered polyol from
the PU and the bottom layer by the excess of glycolysis agent
and reaction by-products [7]. In previous work [8,9], the two-
phase glycolysis of exible PU foams treatment has been
investigated, in relationship with the glycolysis agent and
* Corresponding author. Tel.: 34 926 295300x3416; fax: 34 926 295318.
E-mail address: juan.rromero@uclm.es (J.F. Rodr guez).
0141-3910/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymdegradstab.2007.11.026
Available online at www.sciencedirect.com
Polymer Degradation and Stability 93 (2008) 353e361
www.elsevier.com/locate/polydegstab
the catalysts. The primary aim of these studies was to establish
the inuence of the glycolysing compound on the process and
the quality of the recovered polyol. The works reported the use
of alkaline carboxylates as novel catalysts for the transesteri-
cation of PUs, providing in the presence of diethylene glycol
(DEG), a recovered product suitable to be foamed again after
a slight purication [10].
After the rst approach to the novel polyether polyol recov-
ery process, a detailed study of the main reaction parameters
affecting the reaction and properties of the recovered polyol
has been carried out. It includes carboxylate catalyst concen-
tration, reaction temperature and mass ratio of treated foam
to the glycolysis agent.
2. Experimental
2.1. Materials and methods
Industrial samples of exible PU foam based on polyether
polyol [poly(propylene oxideeblockeethylene oxide) M
w
w3500, functionality with respect to OH groups: 3] and
toluene diisocyanate (TDI) were scrapped with an arbitrary
diameter ranging from 5 to 25 mm. These foams had been pre-
pared in the presence of a cell regulator (surfactant), crosslink-
ing agent, catalyst, colouring agent, mineral loads and water as
a foaming agent. The scrap foam was reacted in several mass
ratios, with diethylene glycol as glycolysis agent (DEG) (PS,
from Panreac, Spain). The ratios of glycolysis agent to PU
foam ranged from 0.75 to 2 by weight. Potassium octoate
was used as catalyst (potassium 2-ethylhexanoate 46.4% by
weight in decyl alcohol (isomers mixture), from NUSA,
Spain).
The glycolysis reactions were carried out in a jacketed 1 L
ask equipped with stirrer and reuxing condenser under
nitrogen atmosphere to avoid oxidation. The glycolysis agent
was placed in the ask and when the temperature raised the
desired level, the required quantity of scrap foam was added
during an hour by means of a continuous feeder, according
to its dissolution. The zero time for the reaction was taken
when all the foam was fed. Temperature was maintained
constant during the feeding and the reaction. Experiments
were carried out in the temperature interval 175e195

C.
2.2. Characterization
At given time intervals aliquots were sampled, cooled and
centrifuged to ensure the total separation of phases. They
were dissolved in tetrahydrofuran (THF from Panreac, Spain)
at a concentration of 1.5 mg mL
1
and then ltered (pore size
0.45 mm). Gel permeation chromatography (GPC) was used to
determine the molecular weight distribution (MWD) as well as
concentration of polyol in the products. Measurements were
performed with a Shimadzu chromatograph (Kyoto, Japan)
equipped with two columns (Styragel HR2 and Styragel
HR0.5) using THF as eluent at 40

C (ow: 1 mL min
1
)
and a refractive index detector. Poly(ethylene glycol) stan-
dards (from Waters, USA) were used for MWD calibration
and mixtures of the industrial starting polyether polyol and
DEG were used as concentration standards. The glycolysis
products were separated and their properties analyzed.
Hydroxyl number and acidity were determined by standard
titration methods (ASTMD-4274-88 and ASTMD-4662-93,
respectively). Amine values in products were determined by
a titration method based on ASTMD-2073-92, and the solvent
was changed for a mixture of 1:1 tolueneeethanol. Water
content was determined by the Karl-Fisher method and the
viscosity was measured by a Brookeld LVTDV-II rotational
viscometer. All chemicals used in these analyses were of the
quality required in the standards. Chemical structures of gly-
colysate products were studied by Fourier Transform Infrared
Spectroscopy using a Perkin Elmer 16PCFT-IR spectrometer;
droplet samples were impregnated on KBr discs.
3. Results and discussion
3.1. Catalyst concentration
As reported in a previous work, the transesterication of
urethanes with the hydroxyl groups of glycols proceeds very
slowly in the absence of a catalyst [9]. This fact has been re-
marked in the literature [11] for this kind of interchange reac-
tions in condensation polymers. Catalysts can reduce the time
required to complete the polyol recovery process, however, the
amount used should be studied, at least due to economic
reasons. The use of potassium octoate as transesterication
catalyst for the PU glycolysis has provided a novel way to
enhance the polyol recovery, and the following step previously
to its implantation comprises a study in deep of main reaction
parameters.
Firstly, degradation reactions of the PU foams were carried
out in the same conditions, just varying the amount of catalyst
added. In Fig. 1 the evolution of urethane oligomer content in
the polyol phase with reaction time is depicted as a function of
0 50 100 150 200 250 300 350
0
5
10
15
20
25
30
%

b
y

w
e
i
g
h
t
reaction time (min)
0 %
1.1%
2.2 %
3.3 %
5.6 %
7.8 %
Fig. 1. Evolution of oligomer content in the upper phase during the glycolysis
reaction of PU foams for different catalyst concentrations, in the presence of
DEG as glycolyzing agent. T
R
a
190

C; W
glycol ag.
/W
PU
1.5.
354 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361
the catalyst amount used. The exchange reaction with DEG,
which occurs randomly according to Casassa theory of
statistical degradation [12], leads to a drastic decrease of PU
molecular weight. The oligourethanes so produced and their
subsequent disappearance give information about the degrada-
tion in that when complete disappearance has been achieved, it
is assumed that all the PU has been degraded. As observed,
a small amount of catalyst induces a strong enhancement of
the degradation, which fully agrees with the common experi-
ence with other condensation polymers like polycarbonate
(PC) [13]. The successive increases in the catalyst concentra-
tion also keep decreasing the reaction time for the complete
degradation, apparently without restriction.
As a result of the PU degradation, the recovered polyether
polyol is released to the reaction medium. As shown in Fig. 2,
polyol release runs parallel to oligomer disappearance. At
given reaction times up to the end point, the more the catalyst
used the more the polyol present in the upper phase due to
faster degradation. Nevertheless, once the plateau value is
achieved, the polyol purity decreases with the catalyst
concentration. In order to gain a better understanding of
the catalyst concentration inuence on the process, in
Fig. 3 the times required to reach the complete PU degradation
and the polyol concentrations obtained in the upper phase for
each concentration studied are presented.
As regards the degradation time, it must be noticed that at
low percentages a slight increase in the catalyst concentration
affects the reaction rate more markedly, as shown by the slope
of the graph of reaction rate versus the catalyst concentration.
The common way to discuss a reaction development is related
to kinetic calculations, but in the literature there are no kinetic
models for PU glycolysis. However, several analogies can be
found with the transesterication processes (glycolysis) to
prepare or degrade common polyesters like PET (poly(ethyl-
ene terephthalate)).These products have gained increasing
commercial importance and therefore there is more literature
dealing with kinetic approaches to describe the inuence of
their transesterication catalysts. Stier et al. [14] have
proposed that the kinetic constant of transesterication of di-
methyl 2,6-naphthalenedicarboxylate (2,6-DMN) with monop-
ropylene glycol (MPG) and metal carboxylates as catalysts can
be expressed as
k k
0
k
0
cat 1
where the constant k
0
takes the uncatalysed reaction into ac-
count and the second term takes care of the catalyzed reaction
assuming that the reaction is of rst order with respect to the
catalyst concentration. Nevertheless, several authors [14e16]
have observed a nonlinear correlation between reactivity and
catalyst concentration. The reaction order is about zero for
a large concentration range and rst order for very low catalyst
concentration.
Experimental results show that glycolysis of PET is analo-
gous to glycolysis of PUs and the kinetic theory used for PET
can also be applied to PU. Firstly, the process can be
conducted in the absence of a catalyst, related to the kinetic
constant k
0
. Secondly, Fig. 3 shows that at low catalyst
concentration, there is a strong dependence with the improve-
ment in the reaction time, but at high concentrations the slope
of the dependence curve decreases. It is expected that at higher
catalyst concentrations the improvement in the reaction rate
would not be noticeable, approaching zero order behaviour.
On the other hand there is a conicting effect on the process
because when the catalyst concentration is increased the nal
concentration of polyol in the upper phase is reduced. This
concentration tends to a minimum steady value at the highest
catalyst concentrations. The GPC chromatograms of upper
phase samples obtained after the complete glycolysis of PU
(Fig. 4) at high and low catalyst concentrations show the
component prole obtained.
In both cases, the main peak corresponds to the recovered
polyol, followed by a group of peaks due to low molecular
weight reaction by-products and that corresponding to DEG.
After the degradation, no oligomers are detected; their peaks
0 50 100 150 200 250 300 350
0
10
20
30
40
50
60
70
80
90
reaction time (min)
%

b
y

w
e
i
g
h
t
0 %
1.1 %
2.2 %
3.3 %
5.6 %
7.7 %
Fig. 2. Evolution of polyol content in the upper phase during the glycolysis
reaction of PU foams for different catalyst concentrations, in the presence
of DEG as glycolyzing agent. T
R
a
190

C; W
glycol ag.
/W
PU
1.5.
0
1 0 2 3 4 5 6 7 8 9
50
100
150
200
250
300
catalyst concentration (%)
r
e
a
c
t
i
o
n

t
i
m
e

(
m
i
n
)
74
76
78
80
82
84
%

p
o
l
y
o
l

b
y

w
e
i
g
h
t
Fig. 3. Dependence of polyol concentration in the upper phase and time to
reach the complete polyol recovery as a function of the catalyst concentration.
T
R
a
190

C; W
glycol ag.
/W
PU
1.5.
355 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361
should appear at retention times lower than the polyol. Al-
though the polyol recovery in the upper phase is quantitative,
a small amount of the bottom phase pollutes the upper phase,
which reveals itself as the low weight compounds group in
the GPC chromatograms. When the catalyst concentration is
increased, not only is the amount of carboxylate salt increased
in the reaction mixture but also the catalyst solvent which is
preferably dissolved in the polyol phase. Moreover the in-
crease of the catalyst concentration enhances the mutual phase
solubility and favours the recovered polyols becoming more
polluted. The properties of the polyol phases obtained, which
are reported in Table 1, are in agreement with the previous
statements: hydroxyl number, amine value and potassium
content, related to solubilisation of DEG and solvent alcohol,
reaction by-products and the catalyst salt, respectively, also
increase with the amount of catalyst employed and so does
the deviation from commercial conformities. Values of proper-
ties out of commercial conformity, especially amine value and
alkaline cations content can negatively affect the further
foaming of the recovered polyol in a new exible PU [17]
and should be reduced to a minimum. An easy, low-cost wash-
ing process allows a good decrease in hydroxyl number but not
a strong one in those undesirable properties [10]. This fact
implies that although the use of a carboxylate catalyst is
strongly recommended due to their improvement in the degra-
dation rate, the concentration used must be reduced in order to
obtain the highest quality of recovered polyol. Probably the
use of 2.2% of catalyst would be the best choice.
3.2. Inuence of temperature
PU foams contain different structures, mainly correspond-
ing to the monomers (polyether polyol) and the groups derived
from the starting isocyanates, like urethane, urea, allophanate
and others. The beginning of thermal degradation of PU
compounds and their derivatives is around 120e250

C. This
range, which is strongly inuenced by the physical character-
istics of the PU, namely, internal crosslinking, hydrogen bonds
and the inner crystalline structure [18], is quite similar to that
described for glycolysis, 180e220

C [19].
Among the groups which form a PU, biuret and allophanate
are the less stable, because their complete thermal degradation
can be reached at 170e180

C [20,21]. Decomposition of
these thermolabile structures yields the precursor groups: ure-
thane, urea, hydroxyl, isocyanate and amine. Urethane and
urea bonds, followed by isocyanurate structures are the follow-
ing in the stability range. It can be necessary to reach 270

C
to achieve their decomposition. Specically, for urethane
groups related to primary and secondary alcohols of a exible
polyether polyol of M
w
3000 g mol
1
, Ravey and Pearce [20]
observed that degradation started at 200

C, whereas Lefebvre
et al. [22] observed that the degradation of urethane groups
started at 250

C for a similar PU. When a rigid polyether pol-
yol is linked to the urethane group, the beginning of the deg-
radation decreases to 230

C [23]. The urethane group thermal
degradation occurs randomly in the polymer structure and has
been described according the following mechanisms [20,24]:
Depolymerisation. Related to the dissociation of the group
in the precursors, polyol and isocyanate. This reaction also
retention time (min)
i
n
t
e
n
s
i
t
y

(
a
.
u
.
)
r
e
c
o
v
e
r
e
d

p
o
l
y
o
l
D
E
G
5.6 %
1.1 %
M
w

4
6
0
M
w


3
0
0
10 12 14 16 18
M
w


1
8
0
M
w

<

M
w
D
E
G
Fig. 4. GPC chromatograms of the upper phases obtained after complete gly-
colysis reactions carried out with several catalyst concentrations. T
a
R
190

C;
W
glycol ag.
/W
PU
1.5; M
w polyol 1.1%
3325, P 1.067; M
w polyol 5.6%
3402,
P 1.059.
Table 1
Properties of recovered polyols (upper phase) obtained with different catalyst concentrations in the glycolysis agent (DEG)
Catalyst concentration
(by weight)
0.0% 1.1% 2.2% 3.3% 5.6% 7.8%
Viscosity 25

C (cP) 838 e 591 600 551 539
Water (%) e 0.12 0.21 0.60 1.03 1.17
Acidity (mg KOH/g) <0.05 <0.05 <0.05 <0.01 <0.01 <0.05
OH number (mg KOH/g) e 161 171 186 195 206
Total amine (mg KOH/g) e e 9.59 11.19 14.41 13.93
Primary amine (mg KOH/g) e e 8.86 10.55 13.41 13.17
K content (ppm) e 3 11 21 44 52
T
R
a
190

C; W
glycol ag.
/W
PU
1.5. Raw polyol conformities: viscosity 560e630 cP, acidity 0.1 mg KOH/g, water content 0.1%, OH number 48 mg KOH/g,
alkalines 5 ppm.
356 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361
takes place for urea groups present in the PU foam. The
resulting products are the starting isocyanate and an aromatic
amine (toluene diamine, TDA). Both reactions are shown in
reactions (2) and (3):
HN
C
O
O R'
R'. HO +
R N C O
Polyurethane Isocyanate Polyol
R
Polyurea Isocyanate Amine
R. H
2
N +
R N C O
HN
C
O
R NH R
Dissociation. By means of a six-membered ring transition,
the urethane group yields an aromatic amine (TDA), carbon
dioxide and a modied polyol, where the hydroxyl end group
is lost in favour of a terminal unsaturation, as expressed in
Eq. (4)
+ CO
2
R NH
2
HN
C
O
O
CH
2
R
+ CH H
2
C R".
CH
H
R"
Polyurethane Unstauration Amine
Both thermal degradation pathways yield undesirable
products like amines, unsaturated polyols, and isocyanates.
All these products should be avoided during the polyol
recovery in order not to pollute the product for its reuse in
new exible PU foam. Accordingly, glycolysis of PU should
be carried out at temperatures lower than degradation temper-
atures, in order to minimize pyrolysis. This fact is especially
important since Ravey and Pearce [20] have shown that a basic
medium promotes pyrolysis, the octoate transesterication
catalyst being a weak base in aqueous media. Taking into
account the variability of decomposition temperatures, depen-
dent on the PU structures, the thermal stability of the PU foam
employed in this work was determined by thermogravimetric
analysis. In Fig. 5 are shown the TGA and DTG curves of
the PU foam under helium atmosphere. The thermograms
show that urethane group degradation starts at 250

C, being
completed at 330

C. The following weight decrease, which
starts at 380

C, corresponds to ether bond thermolysis, ac-
cording to the described interval of degradation temperatures
375e500

C for different polyether polyols [18,20,22,23].
As shown in the thermogram, glycolysis temperatures close
to 250

C would provide unavoidable reactions of pyrolysis in
the urethane and derived groups. Taking into account this fact,
and with the aim of minimizing these reactions, the maximum
glycolysis temperature for the study was set at 200

C. Fig. 6
represents the evolution of urethane oligomer content in the
upper phase during the glycolysis reaction, for different reac-
tion temperatures studied. During the rst hour of the reaction,
at the lowest temperature, the presence of small fragments of
undegraded, and therefore, not analyzable by GPC, PU was
observed in the samples. As these PU fragments are broken
down to smaller oligomers, its concentration in the upper
phase is increased, as the slope of oligomer content curve
reveals. However, oligomer decomposition also takes place
in parallel, so that the oligomer content reaches a maximum;
after that, all the PU portions are small enough to be analyzed
and a global decrease is observed, as in the case of other
temperatures.
As can be observed, temperature is an important factor
affecting urethane degradation: a slight increase in tempera-
ture of only 20

C can reduce the time required to achieve
the complete degradation by more than 3 h. Polyol concentra-
tion history takes place according to oligomers disappearance,
as shown in Fig. 7. The reaction carried out at 175

C provides
0 100 200 300 400 500 600 700
0
20
40
60
80
100
D
G
T

(
a
.
u
.
)
temperature (C)
w
e
i
g
h
t

l
o
s
s

(
%
)
TGA
DTG
Fig. 5. TGA and DTG of the starting PU foam wastes under helium
atmosphere.
0 50 100 150 200 250 300 350 400
0
5
10
15
20
25
30
%

b
y

w
e
i
g
h
t
reaction time (min)
175 C
184 C
190 C
195 C
Fig. 6. Evolution of oligomer content in the upper phase during the glycolysis
reaction of PU foams at different temperatures. Catalyst concentration in the
glycolysis agent 2.2%; W
glycol ag.
/W
PU
1.5.
(2)
(4)
(3)
357 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361
a slow recovery process, as demonstrated by the low percent-
age of polyol in the phase. It should be noticed also that polyol
concentration at the end point (plateau value), is also slightly
modied with reaction temperature.
The inuence of temperature on the polyol concentration in
the recovered product as well as on the time required for com-
plete degradation is summarized in Fig. 8. An apparent linear
relationship between temperature and degradation time can be
observed. The process cannot be properly simplied as the
transesterication of a urethane with a glycol to yield polyol
allowing a simple treatment. Glycolysis of a water-blown
PU comprises multiple secondary reactions and therefore
a whole group of kinetic constants which requires a complex
treatment far from the aim of this article. Nevertheless, the
data obtained provide a helpful tool to optimize the reaction
conditions in a polyol recovery process. If inuences of cata-
lyst and temperature on the process are compared, it is worth
nothing that temperature inuence is more meaningful and
a small detriment can cause a strong delay in the time required
to the complete decomposition. This fact agrees with the
results obtained by Chen et al. [25] for a factorial design of
PET glycolysis, and probes again the similarities between
both glycolysis processes of condensation polymers.
The change in polyol concentration observed is not so
relevant to determine whether secondary reactions are taking
place at the highest temperatures, although the decrease
observed between the extremes of the interval points out to
this fact. If the properties of the recovered polyol obtained
at different temperatures are compared (Table 2), there is
real evidence of the increase of the extent of secondary reac-
tions, which affects the recovery process negatively. Amine
value increases with temperature up to 190

C as a result of
the mentioned secondary reactions. On the other hand
hydroxyl number also increases due to the breakdown of the
transesterication carbamates. Hydroxyl number relates the
amount of hydroxyl end groups of a molecule and its molec-
ular weight in such a way that long molecules like polyol
exhibit a low value. When transesterication takes place in
the urethane group the polyol molecule is substituted by a gly-
col molecule, yielding a non-polymeric low weight carbamate
with higher hydroxyl number. However, if the reaction
conditions become more extreme, the low weight carbamate
is not formed, or is destroyed, yielding an aromatic structure
and glycol. That implies molecules with lower weight and
more hydroxyl groups are not linked; this is revealed as an
increase in the hydroxyl number even if the global level of
polyol pollution remains constant.
Other evidence of the presence of undesirable products in
the recovered polyol is found in the IR spectra of the polyol
phase (Fig. 9). The polyol spectrum has been previously dis-
cussed [8,9], and the presence of impurities can be summa-
rized as: an increase of absorption bands due to stretching
vibrations of hydroxyl end groups around 3460 cm
1
;
presence of bands assigned to stretching vibrations of the
C]O bond for the resulting transesterication carbamates
[9,14,21], found around 1690e1749 cm
1
, linked to a band
at 1520e1540 cm
1
; and presence of a band around
1625 cm
1
assigned to bending vibrations of NH in primary
amines (TDA). As reaction temperature approaches the maxi-
mum, the relative intensity of the bands assigned to impurities
is increased, especially for those assigned to TDA. As
mentioned before, these impurities are strongly undesirable
and can negatively affect the reuse of the recovered polyol
in exible foam applications. Because the lower the reaction
temperature the purer is the polyol recovered, the reaction
0 50 100 150 200 250 300 350 400
0
10
20
30
40
50
60
70
80
90
%

b
y

w
e
i
g
h
t
reaction time (min)
175 C
184 C
190 C
195 C
Fig. 7. Evolution of polyol content in the upper phase during the glycolysis
reaction of PU foams at different temperatures. Catalyst concentration in the
glycolysis agent 2.2%; W
glycol ag.
/W
PU
1.5.
170 175 180 185 190 195 200
0
50
100
150
200
250
300
temperature (C)
r
e
a
c
t
i
o
n

t
i
m
e

(
m
i
n
)
76
78
80
82
84
86
88
90
%

p
o
l
y
o
l

b
y

w
e
i
g
h
t
Fig. 8. Dependence of polyol concentration in the upper phase and time to
reach the complete polyol recovery as a function of temperature. Catalyst con-
centration in the glycolysis agent 2.2%; W
glycol ag.
/W
PU
1.5.
Table 2
Properties of recovered polyols (upper phase) obtained at different reaction
temperatures
Reaction temperature 175

C 184

C 190

C 195

C
Viscosity 25

C (cP) e e 591 694
Acidity (mg KOH/g) 0.012 0.039 0.018 0.015
OH number (mg KOH/g) 158 166 171 172
Total amine (mg KOH/g) 3.78 4.52 9.59 9.80
K content (ppm) 7 e 11 4
Catalyst concentration in the glycolysis agent 2.2%; W
glycol ag.
/W
PU
1.5.
358 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361
temperature should be reduced to the limit. Since proper foam
dissolution was not found at 175

C, the optimal value would
be around 180e185

C.
3.3. Mass ratio of glycolysis agent
In a two-phase glycolysis of PU, the amount of glycol
added is much larger than the stoichiometric quantity
required to produce the transesterication. The excess is
used not only to displace the equilibrium to the glycol substi-
tution, but also to promote the phase splitting which allows the
obtaining of a polyol-rich phase. However, a too large excess
of glycol would imply huge volume equipment requirements
and larger amounts of bottom phase that has to be recycled
by distillation. In order to determine the optimal amount of
glycol a series of degradation experiments was carried out,
where all the reaction conditions were kept constant and the
mass ratio of glycolysis agent to PU foam was modied
from 0.75 to 2. In these experiments the catalyst concentration
in the glycolysis agent also was maintained. Modesti [24] has
described glycolysis experiments even for mass ratios of gly-
colysis agent to PU foam of 0.25, however, when the lowest
amount of glycol was used (ratio 1:0.75) the foam contact
with DEG was very complicated due to the big volume of
the foam scraps fed in comparison with the small volume of
the glycolysis agent added. The foam impregnation was
initially difcult and became even more difcult by the swell-
ing of the non-dissolved portion of PU which made agitation
almost impossible and it was not possible to obtain a proper
phase separation. For the exible foams assayed the limit for
operation was settled as a mass ratio 1:1. Once the amount
of glycol was increased, problems in the foam solution and
phase separation were not observed.
Figs. 10 and 11 present the evolution of oligomer content
and polyol concentration, respectively, in the upper phase for
the different mass ratios which allowed phase separation and
proper foam solution. Both gures show that there are only
slight differences in the process once a minimum mass ratio
value is reached. Oligomers and polyol exhibit, from mass
ratio 1.3, close concentration proles. In contrast, polyol
release in the glycolysis carried out in a 1:1 mass ratio is
delayed, probably due to problems of diffusion of this
polymeric molecule in the swollen cellular PU structure in
the rst steps of the degradation. In this case, not only does
the reaction proceed slower, but also the polyol concentration
obtained is lower.
If dependence on mass ratio of polyol concentration in the
upper phase and reaction time are studied (shown in Fig. 12),
it is observed that changes in the glycolysis agent mass ratio
mainly affect the characteristics of the recovered polyol, in
contrast to the poor inuence exerted on the reaction rate.
Whereas reaction time does not differ too much of those
obtained when the glycolysis agent amount is doubled, polyol
concentration exhibits a steady increase.
When the glycolysis products splitting takes place the
polyol-rich phase is polluted with a small amount of the bottom
phase. This latter is constituted by most of the glycol used and
the reaction by-products, which are preferably dissolved by the
more polar glycol phase. As the amount of glycol is increased
the bottom phase acts as an impurity-extracting compound.
That yields a noticeable net increment of polyol concentration,
which is expected to reach a maximumvalue at very high glycol
amounts, corresponding to the solubility of the glycol (DEG) in
polyol. This value has been found to be 87.5% at room
temperature for the starting polyol [10].
4000 3000 2000 1000
195
190
184
175
wavenumber (cm
-1
)
T

(
a
.
u
.
)

Fig. 9. Comparison of IR spectra of the polyols recovered at different
temperatures.
0 40 80 120 160 200 240 280
0
5
10
15
20
25
30
%

b
y

w
e
i
g
h
t
reaction time (min)
RATIO - 2
RATIO - 1.5
RATIO - 1.3
RATIO -1
Fig. 10. Evolution of oligomer content in the upper phase during the glycolysis
reaction of PU foams for different ratios of glycolysis agent (DEG) to the
treated PU foam. Catalyst concentration in the glycolysis agent 2.2%;
T
R
a
190

C.
359 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361
If chromatograms of the bottom phases obtained are com-
pared (represented in Fig. 13) the dilution effect of the glycol
as a decrease in the relative intensity of the main reaction by-
products peaks (IeIII) in relationship to the obtained for DEG
can be noticed. It is also remarkable that at the lowest mass
ratio, the phase separation problems observed at macroscopic
scale are also revealed as polyol loss in the bottom phase. For
the other mass ratios studied no loss of polyol were detected in
the bottom phase, as revealed by the absence of the polyol
peak in the chromatogram.
Properties of the recovered polyols and phases yields
obtained at the different mass ratios studied are reported in
Table 3. The table also shows that basic impurities are
efciently removed by bigger amounts of glycol and other
properties, like viscosity, are affected by low mass ratios.
But the table also shows the drawbacks of using a large excess
of glycolysis agent: the percentage amount of polyol phase is
decreased. This would affect the economy of the process since
the amount of glycol used is increased as well as the bottom
phase which is a side product to manage. Further studies
[26] have demonstrated that the bottom phase can be success-
fully treated in order to obtain industrially valuable products.
This means that in order to obtain the highest quality
recovered products, the glycolysis must be carried out at
a very large excess of glycol. The limit of this value has to
be established only on the basis of economic criteria.
4. Conclusions
Potassium octoate shows a proper performance as catalyst
in the glycolysis of PU wastes. An increase in the reaction
temperature and catalyst concentration enhances the degrada-
tion rate, however, this also negatively affects the process by
promotion of secondary reactions and contamination of the
polyol phase. As the catalyst concentration is raised, the more
polluted is the polyol obtained. The catalyst of 2.2% seems to
0 40 80 120 160 200 240 280
0
10
20
30
40
50
60
70
80
90
%

b
y

w
e
i
g
h
t
reaction time (min)
RATIO - 2
RATIO - 1.5
RATIO - 1.3
RATIO -1
Fig. 11. Evolution of polyol content in the upper phase during the glycolysis
reaction of PU foams for different ratios of glycolysis agent (DEG) to the
treated PU foam. Catalyst concentration in the glycolysis agent 2.2%;
T
R
a
190

C.
1,0 1,2 1,4 1,6 1,8 2,0
100
120
140
160
180
200
ratio glycolysis agent/PU foam
r
e
a
c
t
i
o
n

t
i
m
e

(
m
i
n
)
76
78
80
82
84
%

p
o
l
y
o
l

b
y

w
e
i
g
h
t
Fig. 12. Dependence of polyol concentration in the upper phase and time to
reach the complete polyol recovery as a function of glycolysis agent (DEG)
to PU foam ratio. Catalyst concentration in the glycolysis agent 2.2%;
T
R
a
190

C.
10 12 14 16 18
retention time (min)
Polyol I
II
III
DEG
RATIO-1
RATIO-1.3
RATIO-1.5
RATIO-2
i
n
t
e
n
s
i
t
y

(
a
.
u
.
)
Fig. 13. GPC chromatograms of bottom phases obtained at different ratios of
glycolysis agent to treated PU foam. Catalyst concentration in the glycolysis
agent 2.2%; T
R
a
190

C; t
R
240 min. Peaks IeIII: reaction by-products.
Table 3
Properties of recovered polyols (upper phase) and phases yield obtained at
different ratios of glycolysis agent to treated PU foam
Mass ratio, glycolysis agent to PU foam 1 1.3 1.5 2
Viscosity 25

C (cP) 681 543 591 569
Water (%) 0.12 0.13 0.10 0.10
Acidity (mg KOH/g) <0.10 <0.10 <0.10 <0.10
OH number (mg KOH/g) 188 188 171 192
Total amine (mg KOH/g) 15.47 10.68 9.59 7.27
Phases yield (%)
Polyol (upper) 36.7 43.7 33.8 28.0
Glycol (bottom) 54.7 50.0 57.4 68.0
Interphase 0.0 4.9 4.5 0.7
Solids 8.6 1.4 4.3 3.3
Catalyst concentration in the glycolysis agent 2.2%; T
R
a
190

C.
360 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361
be the best choice. Temperatures lower than 170

C provide too
slow degradation rates, whereas values higher than 200

C also
promote secondary decarboxylation reactions. In this interval,
an increase in the reaction temperature speeds the degradation
up markedly.
Related to the glycolysis agent amount, the minimum
quantity required to split the phases has been determined, as
well as the optimal ratio. Increasing this ratio does not provide
any additional improvement of the process and affects
economy of the process negatively. As an increment of the
foam/glycol ratio larger than 1:1.5 does not produce a signi-
cant improvement of the process that proportion can be
selected as the best value to carry out the process.
Acknowledgments
Financial support from REPSOL-YPF S.A. through the
General Foundation of the UCLM and CICYT (CTQ2005-
07315) is gratefully acknowledged.
References
[1] Grigat E. Hydrolyse von Kunststoffabfallen. Kunststoffe 1978;68(5):
281e4.
[2] Troev K, Grancharov G, Tsevi R, Tsekova A. A novel approach to recy-
cling of polyurethanes: chemical degradation of exible polyurethane
foams by triethyl phosphate. Polymer 2000;41:7017e22.
[3] Kanaya K, Takahashi S. Decomposition of polyurethane foams by alka-
nolamines. J Appl Polym Sci 1994;5:675e82.
[4] Borda J, Pasztor G, Zsuga M. Glycolysis of polyurethane foams and elas-
tomers. Polym Degrad Stab 2000;68:419e22.
[5] Wu CH, Chang CY, Cheng CM, Huang HC. Glycolysis of waste exible
polyurethane foam. Polym Degrad Stab 2003;80:103e11.
[6] Modesti M, Simioni F. Chemical recycling of reinforced polyurethane
from the automotive industry. Polym Eng Sci 1996;36(17):2173e8.
[7] Hemel S, Held S, Hicks D, Hart M. Chemisches recycling von PUR-
Weichschaumstoffen. Kunststoffe 1998;88(2):223e6.
[8] Molero C, de Lucas A, Rodr guez JF. Recovery of polyols from exible
polyurethane foam by split phase glycolysis: glycol inuence. Polym
Degrad Stab 2006;91:221e8.
[9] Molero C, de Lucas A, Rodr guez JF. Recovery of polyols from exible
polyurethane foam by split phase glycolysis with new catalysts. Polym
Degrad Stab 2006;91:894e901.
[10] Molero C, de Lucas A, Rodr guez JF. Purication by liquid extraction of
recovered polyols. Solvent Extr Ion Exch 2006;24:719e30.
[11] Kricheldorf HR, Denchev Z. Interchange reactions in condensation poly-
mers and their analysis by NMR spectroscopy. In: Fakirov S, editor.
Transreactions in condensation polymers. Weinheim: Wiley-VCH;
1999. p. 1e78.
[12] Cassasa EF. J Polym Sci 1949;4:405e8.
[13] Devaux J. Model studies of transreactions in condensation polymers. In:
Fakirov S, editor. Transreactions in condensation polymers. Weinheim:
Wiley-VCH; 1999. p. 125e58.
[14] Stier U, Gahr F, Oppermann W. Kinetics of transesterication of di-
methyl 2,6-naphthalenedicarboxylate with 1,3-propanediol. J Appl
Polym Sci 2001;80:2039e46.
[15] Tomita K. Studies on the formation of poly(ethylene terephtalate): 1.
Propagation and degradation reactions in the polycondensation of bi(2-
hydroxyethyl) terephtalate. Polymer 1973;14:50e4.
[16] Santacesaria E, Trulli F, Minervini M, di Serio R, Tesser S, Contessa J.
Kinetic and catalytic aspects in melt transesterication of dimethyl
terephthalate with ethylene glycol. J Appl Polym Sci 1994;54:
1371e84.
[17] Woods G. Flexible polyurethane foams: chemistry and technology. Bark-
ing, Essex: Applied Science Publishers; 1982.
[18] Pielichowski K, Kulesza K, Pearce EM. Thermal degradation studies on
rigid polyurethane foams blown with pentane. J Appl Polym Sci
2003;88:2319e30.
[19] Dieterich D, Uhlig K. Polyurethanes. In: Ullmans encyclopedia of indus-
trial chemistry, vol. A21. Weinheim, Germany: VCH; 1992.
[20] Ravey M, Pearce EM. Flexible polyurethane foam. I. Thermal decompo-
sition of a polyether-based, water-blown commercial type of exible
polyurethane foam. J Appl Polym Sci 1997;63:47e74.
[21] Bakirova IN, Vluev VI, Demchenko IG, Zenitova LA. Chemical structure
and molecular mass characteristics of products in the glycolysis of a ex-
ible poly(urethane) foam. J Polym Sci A 2002;44(6):615e22.
[22] Lefebvre J, Bastin B, le Bras M, Duquesne S, Paleja R, Delobel R. Ther-
mal stability and re properties of conventional exible polyurethane
foam formulations. Polym Degrad Stab 2005;88:28e34.
[23] Lefebvre J, Duquesne S, Mamleev V, le Bras M, Delobel R. Study of
the kinetics of pyrolysis of a rigid polyurethane foam: use of the invari-
ant kinetics parameters method. Polym Adv Technol 2003;14:
796e801.
[24] Modesti M. Recycling of polyurethane polymers. In: Advances in ure-
thane science and technology, vol. 13. Lancaster: Tecnomic Publishing
Co.; 1996. 237e288.
[25] Chen CH, Chen CY, Lo YW, Mao CF, Liao WT. Studies of glycolysis
of poly(ethylene terephtalate) recycled from postconsumer soft-drink
bottles. II. Factorial experimental design. J Appl Polym Sci 2001;80:
956e62.
[26] Molero C, de Lucas A, Rodr guez JF. Utilization of by-products origi-
nated in the chemical recycling of exible polyurethane foams. In:
Muller-Hagedorn M, Bockhorn H, editors. Feedstock recycling of plas-
tics. Karlsruhe: Universitatsverlag Karlsruhe; 2005. p. 231e7.
361 C. Molero et al. / Polymer Degradation and Stability 93 (2008) 353e361

Potrebbero piacerti anche