You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/327521596

Isocyanurate formation during rigid polyurethane foam assembly: A


mechanistic study based on: In situ IR and NMR spectroscopy

Article  in  Polymer Chemistry · September 2018


DOI: 10.1039/C8PY00637G

CITATIONS READS

18 1,241

5 authors, including:

Abdulghani Al Nabulsi Daniela Cozzula


RWTH Aachen University Birkenstock Production
2 PUBLICATIONS   18 CITATIONS    32 PUBLICATIONS   504 CITATIONS   

SEE PROFILE SEE PROFILE

Walter Leitner Thomas Ernst Müller


RWTH Aachen University Ruhr-Universität Bochum
568 PUBLICATIONS   22,900 CITATIONS    178 PUBLICATIONS   9,531 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Dream Production View project

Bioressources View project

All content following this page was uploaded by Daniela Cozzula on 24 October 2018.

The user has requested enhancement of the downloaded file.


Polymer
Chemistry
PAPER

Isocyanurate formation during rigid polyurethane


Cite this: Polym. Chem., 2018, 9,
foam assembly: a mechanistic study based on
4891 in situ IR and NMR spectroscopy†
A. Al Nabulsi,a D. Cozzula, *a T. Hagen,b W. Leitner c
and T. E. Müller *‡c

An in-depth exploratory study on the mechanism of isocyanurate formation during the assembly of rigid
PU/PIR polyurethane foams was performed thereby unravelling the role and working principle of the cata-
lyst. Mimicking the complex pattern of the foam with mono-functional equivalents, the pathways to iso-
cyanurate formation were investigated. Reactions focusing on potassium acetate as benchmark catalyst
were thereto followed by in situ IR and NMR spectroscopy. Two pathways to isocyanurate formation from
isocyanates were revealed: a one-component route via cyclotrimerization and a two-component route
via carbamate and allophanate. We show that in the presence of alcohol and along with the formation of
carbamate, isocyanurate is formed preferentially according to the two-component route in three con-
Received 26th April 2018, secutive steps. Allophanate, which in this study was isolated and characterized, is formed in situ in small
Accepted 31st August 2018
transient concentrations from carbamate and the excess of isocyanate. For the first time it is proven that
DOI: 10.1039/c8py00637g the allophanate intermediate undergoes an addition–elimination step with nucleophilic species to provide
rsc.li/polymers the isocyanurate.

Introduction the material flexible and ductile. A high content of isocyanu-


rate moieties provides a highly cross-linked dense network
Polyurethane–polyisocyanurate (PU/PIR) foams play a substan- that renders the material harder, yet on occasion brittle. In
tial role in the construction market. Their efficiency as thermal controlling these structure–property relationships, the choice
insulating materials, a trait that contributes markedly to con- of the catalyst and the mechanism of isocyanurate formation
serving energy, leads to rapidly increasing relevance for redu- play a key role.
cing the consumption of fossil carbon resources. The highly The general mechanism for the anionic trimerization of iso-
crosslinked isocyanurate structure of the rigid PU/PIR foam cyanates is widely accepted. It was proposed first by Reymore
leads to excellent stability and high mechanical strength. The et al.11,12 and was illustrated later by Hoffman, who prepared
attractivity in construction stems from the distinctive combi- triphenyl isocyanurate from phenyl isocyanate using potass-
nation with a lower density and weight compared to many ium acetate as catalyst.4 According to this mechanism, the
other conventional materials.1–6 This is in addition to the low nucleophile acetate adds to an isocyanate molecule to give a
thermal conductivity and excellent flame retardant properties stable intermediate anion that, in the presence of excess iso-
of rigid PU/PIR foams.1,2,6–10 Strikingly these properties can be cyanate, undergoes intramolecular cyclization thereby regener-
adjusted in a wide range by controlling the ratio of urethane to ating the acetate catalyst.11
isocyanurate moieties. A high content of urethane moieties Alternative to the cyclotrimerization of isocyanate, the iso-
provides a large proportion of linear connections that render cyanurate is also formed through the stepwise reaction
between alcohol and excess isocyanate. Isocyanates readily
react with compounds comprising so-called “active hydrogen
a atoms” to the corresponding carbamoyl derivatives.13–19 In
CAT Catalytic Center, RWTH Aachen University, Worringerweg 2, 52074 Aachen, DE,
Germany. E-mail: Daniela.Cozzula@catalyticcenter.rwth-aachen.de case of alcohol (R′OH) and isocyanate (RNCO), carbamate
b
COVESTRO Deutschland AG, Kaiser-Wilhelm-Allee 60, 51373 Leverkusen, DE, (RNHCOOR′) is formed in the first step. The latter is trans-
Germany formed with excess isocyanate into the corresponding allopha-
c
ITMC, RWTH Aachen University, Worringerweg 1, 52074 Aachen, DE, Germany. nate (RNHCONRCOOR′) that is converted into the stable iso-
E-mail: Thomas.Ernst.Mueller@itmc.rwth-aachen.de
cyanurate thereby regenerating the starting alcohol.13,14,16,18
† Electronic supplementary information (ESI) available. See DOI: 10.1039/
c8py00637g Despite a number of studies on the chemistry of isocyanurates,
‡ Current address: RFH Köln, Schaevenstraße 1 a-b, 50676 Köln, Germany (DE). the details of the illustrated general mechanism have not been

This journal is © The Royal Society of Chemistry 2018 Polym. Chem., 2018, 9, 4891–4899 | 4891
Paper Polymer Chemistry

outlined. In addition, the role of intermediates in the complex


reaction network of PU/PIR foams has remained unclear. By
investigating the mechanism of isocyanate trimerization with
modern spectroscopic techniques, we targeted on the one
hand at generating a fundamental understanding of catalysis
in polyurethane material manufacture, and on the other hand
to set the base for the development of more efficient catalysts.
At present new catalysts are put forward for controlling the
ratio of urethane to isocyanurate moieties.4,20–22 Nevertheless
potassium acetate remains to be the standard catalyst in the
Scheme 1 Mimic reaction chosen to investigate the mechanism of iso-
industrial scale production of PU/PIR foams.
cyanurate formation during rigid polyurethane foam assembly.
The target of this study was to explore in-depth the reaction
network that occurs during the formation of rigid PU/PIR
foams. In that way, we aimed to unravel the role of reaction
tion was performed in a steel reactor equipped with a highly
intermediates and understand the working principles of the
efficient cooling system. Propylene carbonate was selected as
benchmark catalyst potassium acetate. Making use of mono-
solvent due to the fact that the polarity resembles the settings
functional reagents to mimic the complex features of the
in the foam, for its high boiling point and as in IR spec-
foam, the reactions were followed in situ by IR or NMR spec-
troscopy the bands do not interfere with the region of interest.
troscopy. To gain more detailed insight into the reaction
After injecting a solution of acetate (0.1 mol% with respect to
network the nucleophilicity of the catalyst, the ratio of isocya-
pTI) and DEME into a solution of pTI, the intensity of the
nate/alcohol and the catalyst concentration were varied.
characteristic signal of the asymmetric NvCvO stretch of pTI
at 2274 cm−1 began to decrease rapidly (Fig. 1), while two
other signals rapidly increased in the characteristic region for
Results and discussion the CvO stretch. Based on reported data,31 the signal at
Mimic reaction 1728 cm−1 is attributed to the CvO stretch of the carbamate
product (CRB) and the signal at 1704 cm−1 is attributed to the
In the system of PU/PIR foams, the polymer network is gener-
CvO stretch of isocyanurate (ISR).32
ated by reacting polyisocyanates such as methylene diphenyl
In the presence of KOAc, 90% of pTI was converted within
diisocyanate and its higher oligomers (MDI) with polyols.
of 1.2 min (Fig. 1, top) at an initial rate of 2.50 × 10−3 L mol−1 s−1.
Additives tailor the properties of the PU/PIR foams to the
In the presence of KOAcF, more time (6 min) was required to
specific application. For investigating the mechanism of iso-
obtain the same conversion of pTI (Fig. 1, bottom). The initial
cyanurate formation, we decided to mimic the complex reac-
rate of isocyanate consumption was 1.28 × 10−3 L mol−1 s−1. The
tion system of the PU/PIR foam with related mono-functional
lower activity of KOAcF can be related to the lower nucleophilicity
substrates thereby facilitating the analysis of the spectra as well
of the carboxylate group. In both cases the two products of the
as the reaction products. p-Tolyl isocyanate (pTI) and diethylene
reaction, CRB and ISR, were obtained in the same molar ratio of
glycol methyl ether (DEME) were selected as mimic substrates
1.2 : 1. As the reaction was performed with 2.5 fold excess of pTI
for MDI and ethylene glycol units, respectively, that are com-
(pTI : DEME 3.5 : 1), this means that the isocyanate (28.6%)
monly present in commercial polyether polyols used in PU/PIR
reacted preferentially with the alcohol to carbamate and that only
formulations. Using potassium acetate as nucleophilic catalyst,
the remaining isocyanate (71.4%) trimerized to isocyanurate
two main products are expected: 2-(2-methoxyethoxy) ethyl
(Scheme 1). As the formation of carbamate prevails over isocyanu-
p-tolyl carbamate (CRB) as product of the addition reaction
rate, yet high isocyanurate content is essential for excellent pro-
between one alcohol molecule and one isocyanate molecule,
perties of rigid PU/PIR foams, the understanding of the pathways
and tri-p-tolyl-s-triazinane-2,4,6-trione (ISR) as product of the tri-
to isocyanurate formation is of great importance. It also opens the
merization reaction of three isocyanate molecules (Scheme 1).
prospect to developing catalysts that would be more selective
Kinetic study towards isocyanurate formation.

To obtain insight into the reaction kinetics and the influence


of the nucleophilicity of the catalyst, the catalytic activity of Reaction pathways
potassium acetate (KOAc) and potassium trifluoroacetate In the literature,11 two competitive reaction pathways are
(KOAcF) was compared. Mimicking the conditions of real foam described to explain the formation of isocyanurate: via the
formulations, the reaction was performed at a fixed molar cyclotrimerization of isocyanate (Pathway A); and via three con-
ratio of pTI : DEME (3.5 : 1) at 70 °C. Owing to its well resolved secutive steps involving alcohol (Pathway B) where: (i) alcohol
spectral features, in situ IR spectroscopy was selected to reacts with isocyanate to form carbamate, (ii) carbamate reacts
monitor the components that take part in the reaction.23–25 To with another molecule of isocyanate to form allophanate and
ensure isothermal conditions despite the high exothermicity (iii) allophanate reacts with a third molecule of isocyanate to
of the reaction between isocyanate and alcohol,26–30 the reac- form isocyanurate thereby liberating alcohol (Scheme 2).

4892 | Polym. Chem., 2018, 9, 4891–4899 This journal is © The Royal Society of Chemistry 2018
Polymer Chemistry Paper

Trimerization in the absence of alcohol


The reaction pathway of isocyanurate formation in the absence
of alcohol was investigated first. The trimerization of pTI with
KOAcF (stoichiometric ratio of 1 : 1) was followed with in situ
NMR spectroscopy. KOAcF was selected as catalyst as its lower
activity compared to KOAc allows a more precise recording of
spectra and, thus, more facile monitoring of the reaction com-
ponents. The experiment was carried out in deuterated
dimethyl sulfoxide (DMSO-d6) at 25 °C and the spectra were
recorded every two minutes. pTI was identified by the signal of
the methyl group at the aromatic ring at 2.28 ppm. The
formation of isocyanurate (ISR) was followed by the methyl
group of the aromatic rings at 2.34 ppm. The changes in the
intensity and the chemical shift of these characteristic signals
were followed over 19.5 h. All chemical shifts were referenced
to the shift of DMSO-d6 at 2.50 ppm. The most significant
spectra are shown in Fig. 2 concentrating for clarity at the
beginning and the end of the reaction.
The high-intensity signal at 2.28 ppm (85.5 mol%) in the
spectrum recorded at the early stage of the reaction (2 min) is
attributed readily to pTI. A low intensity signal assigned to ISR
emerged at 2.34 ppm (1.8 mol%). Noteworthy, three new
signals with low intensity were visible in the methyl region at
2.22 (signal A), 2.24 (signal B) and 2.36 ppm (signal C). The
spectrum recorded at 4 min showed a decrease in the signal
intensity of pTI (83.4 mol%) and a corresponding increase in
the signal intensity of ISR (2.9 mol%). Also, the intensities of
the three signals A to C increased with time. The spectrum
recorded at 8 min revealed significant changes in the methyl
region: a considerable decrease in the signal intensity of pTI
(75.6 mol%), a correspondingly strong increase in the signal

Fig. 1 Time-resolved in situ IR spectra showing the consumption of


p-tolylisocyanate ( pTI) when reacted with diethylene glycol methyl
ether (DEME) using KOAc as catalyst (top) in comparison to KOAcF
(bottom).

Fig. 2 Selected 1H-NMR spectra of the reaction mixture of pTI and


Scheme 2 Competitive reaction pathways to the formation of isocya- KOAcF recorded over 19.5 h. For clarity, only the region covering chemi-
nurates from the reaction of alcohols with excess isocyanate (Ar = cal shifts between 2.16 and 2.38 ppm is shown. pTI ( p-tolylisocyanate);
RC6H4). ISR (isocyanurate); urea (Ar–NHC(O)NH–Ar).

This journal is © The Royal Society of Chemistry 2018 Polym. Chem., 2018, 9, 4891–4899 | 4893
Paper Polymer Chemistry

intensity of ISR (8.1 mol%) and an increase in the signal inten- To learn, if signals A to C belong to one of the three
sity of A with a small shift to 2.20 ppm (5.8 mol%). The signals expected intermediates, the in situ NMR spectra of the reaction
B and C also shifted slightly to 2.23 and 2.35 ppm. At 14 min between pTI and KOAcF were analysed in further detail. Note
(not shown for clarity), there was a continued decrease in the that the intensity of signal A steadily increased in intensity
signal intensity of pTI (62.2 mol%) and an increase in the during the reaction. Consistent with the initial ratio of isocya-
signal intensity of ISR (17.6 mol%). The other three signals nate to acetate, even the highest intensity of signal A in the
slightly shifted: signal A at 2.20 ppm shifted to 2.19 ppm, spectrum recorded at 19.5 h was low compared the starting
signal B overlapped with the signal of a substituted urea materials and products. Likewise, the intensity of signals B
(Ar–NHC(O)NH–Ar) at 2.23 ppm and signal C shifted to and C was low passing through a shallow maximum shortly
2.35 ppm and partially overlapped with the signal of ISR. The after the start of the reaction. Close inspection of the aromatic
formation of the substituted urea is attributed to the reaction region of the 1H-NMR spectrum recorded at 19.5 h revealed
of pTI with traces of water in the DMSO-d6. At 24 min, the two doublets at 7.23 and 6.88 ppm that are part of an AA′ BB′
intensity of the pTI signal was decreased further (42.7 mol%), system (Fig. 3). Closer inspection of the recorded spectra
while the intensities of the characteristic signals of ISR revealed that the chemical shift of both doublets shifted
(30.8 mol%) and signal A (13.2 mol%) were increased. The during the reaction, while the same difference in frequency
signals B and C were no longer distinguishable in the spectra (213 Hz) was maintained. The singlet at 2.18 ppm (signal A) is
due to spectral overlap with other bands. After the projected characteristic of the methyl protons of a tolyl substituent. The
transition through the spectrum at 120 min, the spectrum three signals have relative integrals of 2 : 2 : 3. Positions and
recorded at 1170 min (19.5 h) showed only two signals in the integrals match the ones observed for the isolated intermedi-
methyl region in addition to the urea signal: ISR at 2.34 ppm ate I1 (see ESI†). Likewise, we tentatively assign the singlets at
(47.2 mol%) and signal A (30.4 mol%) at 2.18 ppm. 2.24 (signal B) and 2.36 ppm (signal C) to the methyl protons
According to the general mechanism for the anionic trimer- of the distal tolyl substituent of I2 and I3, respectively.
ization of isocyanates that was proposed by Reymore and illus- To verify the structure of intermediate I1, a sample of the
trated by Hoffman4,11 the formation of isocyanurate goes reaction mixture was analysed by ESI-MS spectroscopy.33–36
through three anionic intermediates (Scheme 3): the nucleo- The signals with high abundance in the positive-ion mode are
philic anion adds to one isocyanate molecule to produce a first listed in Table 1. Most importantly, the ion mass signal at
intermediate anion (I1). Then, the addition of this intermedi- m/z = 279.97 with relative abundance of 62% is assigned to
ate to another isocyanate molecule produces a second inter- 2,2,2-trifluoroacetic p-tolylcarbamic anhydride associated with
mediate (I2) that adds to a yet another isocyanate molecule to an additional methanol molecule. This species is attributed to
produce the third intermediate (I3). During a subsequent ring- the addition of CF3COO− to pTI, and the methanol is due to
closure step the latter forms the isocyanurate thereby releasing the use of methanol as medium for the ESI-MS analysis. The
the anion. If one of the steps is rate determining, one specific corresponding ion mass of intermediate I1 after losing one
intermediate is expected to be observed. If similar reaction molecule of carbon dioxide was also observed (m/z = 242.0).
rates were present for the three steps, three intermediates are The ion mass at m/z = 438.12 is assigned to ISR. The corres-
expected to be observed.

Fig. 3 Selected 1H-NMR spectra of the reaction mixture of pTI and


KOAcF recorded over 19.5 h. For clarity, only the aromatic region
Scheme 3 Established catalytic cycle of the isocyanurate formation via between 6.78 and 7.45 ppm is shown. pTI ( p-tolylisocyanate); ISR (iso-
the anionic trimerization of isocyanate ( pathway A). cyanurate); urea (Ar–NHC(O)NH–Ar).

4894 | Polym. Chem., 2018, 9, 4891–4899 This journal is © The Royal Society of Chemistry 2018
Polymer Chemistry Paper

Table 1 High abundance signals observed from the positive-ion mode which continued to accumulate until the complete conversion
of ESI-MS analysis of the reaction mixture of pTI and KOAcF at 19.5 h of pTI.

m/z Relative abundance [%] Assignment Trimerization in the presence of alcohol


190.91 61 (CF3COOK) K+ The reaction pathway of isocyanurate formation in the pres-
279.97 62
ence of alcohol was investigated thereafter. The 1H-NMR study
was repeated (vide supra), whereby the reaction was performed
438.12 50 (Isocyanurate) K+ in the presence of DEME at a molar ratio of 3.5 : 1 : 0.002
494.81 100 (CF3COOK)3 K+ ( pTI : DEME : KOAcF). Along with the characteristic signals of
798.71 12 (CF3COOK)5 K+
pTI and ISR, CRB and the corresponding allophanate (2-(2-
methoxyethoxy) ethyl p-tolyl ( p-tolylcarbamoyl) carbamate,
APH) were identified by the resonance of their –COOCH2
ponding ion masses of ISR plus an additional water molecule
group at 4.18 and 4.23 ppm, respectively. The alcohol (DEME)
(m/z = 456.83) and an additional molecule of CF3COOK (m/z =
was followed by the –OH signal that, in our case, was observed
590.07) were also observed. The ion masses at m/z = 190.91,
at 4.57 ppm.
494.81 and 798.81 are readily assigned to [(CF3COOK)n K]+
The 1H-NMR spectrum at 1 min showed high-intensity
with n = 1, 3 and 5, respectively. The formation of singly
signals at 2.27 ppm (76.8 mol%) and 4.57 ppm (21.9 mol%),
charged ions of [(CF3COOM)n M]+ and [(CF3COOM)n CF3COO]−
attributed to pTI and DEME in the initial ratio (Fig. 5). At
in both positive- and negative ionization modes was reported
37 min, a high-intensity signal at 4.18 ppm (21.3 mol%) as
by Moini et al.37 In that study, alkali–metal trifluoroacetates
well as a low-intensity signal at 4.23 ppm (3.9 mol%) attributed
had been used as calibration standards for ESI-MS. The nega-
to the –COOCH2 groups of the formed CRB and APH were
tive-ion mode provided the corresponding negative ions for
observed. In parallel, the intensity of the methyl signal of pTI
intermediate I1, ISR and [(CF3COOK)n CF3COO]− with n = 1, 3
(62.7 mol%) and the –OH signal of DEME (8.6 mol%)
and 5 (see ESI†). Thus, the ESI-MS data confirm the formation
decreased. In the spectrum at 79 min, there was a considerable
of intermediate I1 as intermediate during the anionic trimeri-
decrease in the intensity of the signals of pTI (49.2 mol%). The
zation of isocyanate and at the same time provide additional
spectrum revealed a small increase in the signal intensities of
evidence for the attribution of signal A to this intermediate.
CRB and APH (26.3 and 7.5 mol%) and the appearance of the
To obtain a reaction profile, the molar ratio was calculated
characteristic methyl signal of ISR (8.1 mol%). At 118 min, the
for each component (mol%) from the normalized integral of
spectrum showed a decrease in the signal intensities of pTI
the corresponding characteristic signal to the sum of all nor-
(40.3 mol%) and APH (6.1 mol%), while a small increase in
malized integrals (see ESI†) and plotted against time. The
the signal intensity of CRB (28.4 mol%) and a considerable
profile (Fig. 4) reveals two phases: a first phase commencing
increase in the signal intensity of ISR (14.9 mol%) were
from mixing the reagents to about 120 min and a second
observed. In the spectrum recorded at 18.3 h, the pTI signal
phase ranging from 120 min to full conversion of isocyanate.
During the first phase, isocyanate was consumed at high rate
according to a pseudo-first order reaction. Trimer and species
I1 formed as main products, reaching a maximum concen-
tration at 120 min. In the second phase, pTI was consumed at
much lower rate. Trimer remained unchanged, while the
molar ratio of signal A increased slightly towards the end of
the reaction. It was also observed that the decrease in the
molar ratio of pTI was equal to the increase in species I1,

Fig. 5 Selected 1H-NMR spectra of the reaction mixture of pTI and


DEME in the presence of the catalyst KOAcF recorded over 18.3 h. For
clarity, the aromatic region is omitted. pTI ( p-tolylisocyanate); CBR (car-
Fig. 4 Reaction profile obtained for the reaction of p-tolyl isocyanate bamate), ISR (isocyanurate), DEME (diethylene glycol methyl ether), APH
( pTI) with KOAcF to isocyanurate (ISR) and species I1. (allophanate).

This journal is © The Royal Society of Chemistry 2018 Polym. Chem., 2018, 9, 4891–4899 | 4895
Paper Polymer Chemistry

was no longer visible and the intensity of the signal of DEME the formation of isocyanurate. Formation of ISR commences,
was low (5.0 mol%). The intensities of the signal of CRB and when APH reaches a critical concentration of 8% molar ratio.
ISR were high (50.5 and 35.6 mol%, respectively). At 4.23 ppm, Thus, isocyanurate is a consecutive product in the further reac-
no signal was detected indicating the complete consumption tion with allophanate rendering allophanate a key intermedi-
of allophanate (Fig. 5). ate in the formation of isocyanurate, when alcohol is present.
To confirm the formation of allophanate during the experi- Consistent with the allophanate phase, ISR formation takes
ment, a sample taken from the reaction mixture was analysed place during the third phase, but only, when APH is present in
by HPLC-MS. Clearly an ion mass at m/z = 387 was observed the reaction mixture.
attributed to the presence of APH in the reaction mixture.
A sample of pure APH was thereafter obtained as solution in Allophanate as key intermediate
methanol by preparative HPLC chromatography and analysed During the reaction between pTI and DEME the formation of
in detail by 1H-NMR spectroscopy and COSY (see ESI†). APH clearly reached a maximum before the phase of isocyanu-
The obtained reaction profile, shown in Fig. 6, reveals three rate formation. In order to determine whether the allophanate
phases preceding the full conversion of isocyanate and the for- was formed from carbamate in a consecutive reaction with iso-
mation of isocyanurate: (i) a first phase from mixing of the cyanate, another experiment was carried out: the alcohol was
reagents to 37 min (carbamate phase), (ii) a second phase introduced as carbamate and the reaction with isocyanate per-
from the 37 min to around 2 h (allophanate phase) and (iii) a formed at a ratio pTI : CRB of 1 : 1 in the presence of stoichio-
third phase from about 2 h to the full conversion of isocyanate metric KOAcF. Due to the higher concentration of KOAcF, the
(isocyanurate phase). rate of the reaction was expected to be high. Therefore, we
During the carbamate phase (Phase i), pTI and DEME focused on the spectra at the early stage (1, 3, 6 and 12 min)
reacted at high rates and the corresponding carbamate (CRB) and final stage (56 min and 20 h, Fig. 7) of the reaction.
formed as main product following roughly a second-order reac- At 1 min, the 1H-NMR spectrum showed high-intensity
tion. A small amount of APH was also produced. ISR formation signals at 2.27 and 4.18 ppm, attributed to pTI (41.7 mol%)
was not observed at this stage of the reaction. In the allopha- and CRB (51.7 mol%), respectively. A low-intensity signal at
nate phase (Phase ii), pTI and DEME were consumed albeit at 4.23 ppm for APH (4.1 mol%) was also detected. Already at
a lower rate, and consequently the formation of CRB was 3 min, the intensities of the characteristic methyl signal of pTI
slower. The concentration of APH increased further during (29.5 mol%) and –COOCH2 signal of CRB (45.3 mol%) rapidly
this phase reaching a maximum at 79 min. This was decreased, while there was a significant increase in the inten-
accompanied by considerable formation of ISR. In the isocya- sity of the characteristic –COOCH2 signal of APH (13.8 mol%).
nurate phase (Phase iii), pTI was consumed together with APH At this stage, both the characteristic methyl signal of ISR
until complete conversion was obtained, while the concen- (6.1 mol%) at 2.34 ppm and the –OH group of DEME
tration of DEME remained almost unchanged. In parallel, CRB (2.4 mol%) at 4.67 ppm were observed. Note that DEME was
and ISR were formed steadily to reach a maximum concen- the corresponding alcohol of the carbamate. The signal inten-
tration at the end of the reaction. During the carbamate phase, sity of pTI (21.4 mol%) continued to decrease in the spectrum
pTI obviously reacted with DEME to the corresponding CRB.
An excess of pTI and a considerable concentration of CRB are
required to achieve the formation of APH as a consecutive
product. The allophanate phase seems to be the critical step in

Fig. 6 Reaction profile recorded for the mixture of pTI and DEME in the Fig. 7 Selected 1H-NMR spectra of the reaction mixture of pTI and CRB
presence of the catalyst KOAcF. pTI ( p-tolylisocyanate), CBR (carbamate), in the presence of KOAcF recorded over 20 h. For clarity, the aromatic
ISR (isocyanurate), DEME (diethylene glycol methyl ether), APH region is omitted. pTI ( p-tolylisocyanate); CBR (carbamate), ISR (isocya-
(allophanate). nurate), DEME (diethylene glycol methyl ether), APH (allophanate).

4896 | Polym. Chem., 2018, 9, 4891–4899 This journal is © The Royal Society of Chemistry 2018
Polymer Chemistry Paper

recorded at 6 min, while the signal intensity of CRB increased


(47.4 mol%). In comparison to the previous spectrum, APH
showed a slight decrease in the signal intensity (13.2 mol%).
On the other hand, an increase in the signal intensity of ISR
(10.3 mol%) accompanied with an increase in the –OH signal
of DEME (5.1 mol%) was observed. While the intensity of pTI
and APH signals had steadily decreased at 6 min, the intensity
of CRB, ISR and DEME signals had continued to increase. The
signals of pTI (2.6 mol%) had a very low intensity at 56 min; so
did the signal of APH (1.5 mol%). The signal of CRB
(57.3 mol%), ISR (19.4 mol%) and DEME (13.2 mol%) had
higher intensities. Both pTI and APH signals were no longer
observed in the spectra recorded afterwards. Signals of
CRB and ISR showed a slight increase (59.4 and 22.1 mol%, Scheme 4 Feasible routes for the formation of ISR (isocyanurate) from
respectively). At 20 h, no further changes were observed in the APH (allophanate).
spectra (Fig. 7).
The reaction profile (Fig. 8) clearly reveals the formation
and consumption of allophanate. After a first phase from • Reaction of I1 with the carbonyl group in α position leads
mixing of the reagents to 3 min (allophanate phase), a second to the formation of carbamate (CRB) and intermediate I2.
phase to the full conversion of isocyanate (isocyanurate phase) Since the generation of uretdione by an intramolecular
is observed. During the first phase, pTI and CRB were con- mechanism of I2 was not observed, the reaction continues and
sumed at high rates. APH was formed as the main product, I2 reacts further with another molecule of pTI to form inter-
reaching a maximum concentration at 3 min. A significant mediate I3. The latter undergoes an intramolecular cyclization.
amount of ISR was also formed. In the second phase, pTI and Thereby, the lone electron pair of the terminal nitrogen moiety
APH were consumed at lower rates, while CRB showed an unex- is added to the carbonyl group γ and the catalytic nucleophile
pected increase. ISR was formed during the second phase departs with an electron pair to form the energetically
until complete conversion of pTI had been obtained. Although favoured six-membered ring (ISR).
CRB replaced DEME as the starting material, the formation of • Reaction of I1 with the carbonyl group in β position leads
DEME was clearly observed accompanied by the formation of to the release of alcohol (DEME) and formation of intermedi-
ISR. Thus, allophanate is a consecutive product formed by the ate I3 that eventually produces the isocyanurate (ISR).
reaction of carbamate and isocyanate. In the second phase, the
formation of isocyanurate is accompanied by the (unforeseen) Parallel pathways to isocyanurate formation
formation of carbamate and alcohol. Two competitive pathways in the reaction of isocyanate to iso-
These observations are readily explained by two analogous cyanurate were identified (Scheme 5):
mechanistic routes (Scheme 4), both of which lead to the for- Pathway A: Cyclotrimerization of isocyanate to isocyanurate
mation of isocyanurate: nucleophilic addition of the inter- (anionic route), where the nucleophilic anion adds to a first
mediate I1 to the carbonyl group in α or β position of the allo- molecule of isocyanate to form the intermediate I1. The latter
phanate molecule accompanied by elimination of carbamate reacts further with two further molecules of isocyanate to form
or alcohol, respectively. intermediates I2 and I3, respectively. The isocyanurate ring is
then formed via an intramolecular nucleophilic substitution,
by which the nucleophilic anion departs for a new catalytic
cycle.
Pathway B: The formation of isocyanurate via three consecu-
tive steps, in which the allophanate plays a key role as inter-
mediate (allophanate route). In this pathway, a molecule of iso-
cyanate adds to a molecule of alcohol to form the corres-
ponding carbamate. Another molecule of isocyanate adds to
the carbamate to form the corresponding allophanate. The
third step consists of nucleophilic addition of the intermediate
I1 to the formed allophanate following one of two analogous
routes that lead both to the formation of the isocyanurate ring
and elimination of either carbamate or alcohol. To promote
the formation of isocyanurate in this step, a significant con-
Fig. 8 Reaction profile of a mixture of pTI and CRB in the presence of
centration of the allophanate is required to be able to react
KOAcF. pTI ( p-tolylisocyanate); CBR (carbamate), ISR (isocyanurate), with the intermediate I1 that is formed in situ in a catalytic
DEME (diethylene glycol methyl ether), APH (allophanate). amount. Also with regard to the preferred reaction pathway, a

This journal is © The Royal Society of Chemistry 2018 Polym. Chem., 2018, 9, 4891–4899 | 4897
Paper Polymer Chemistry

is also reported in literature).38 The rapidly increasing viscosity


gives rise to diffusion limitations that prevent complete con-
version of the isocyanate moieties. Despite closer inspection of
the spectra, the bands of intermediates were obscured by the
intense bands of the main products.

Conclusions
In summary, we have explored in-depth the reaction network
of isocyanurate formation during the reaction of isocyanates
and alcohols. The study has unravelled the fundamental role
of a nucleophilic catalyst as simple as potassium acetate and
resolved a detailed reaction mechanism. Thereby the pro-
perties of the catalyst have to be well balanced. As a strong
nucleophile the catalyst is more prone to bind the isocyanate
molecule while, at the same time, it has to be a good leaving
group to facilitate the ring closure step and, thus, formation of
isocyanurate. In the catalytic cycle, in situ formed species I1 plays
a key role; its importance as longer-living intermediate was
demonstrated. In the presence of alcohol, the reaction proceeds
through a preferred mechanism via carbamate and allophanate
followed by ring closure to isocyanurate. Allophanate was
demonstrated to be a consecutive product of the reaction
between the carbamate and the isocyanate. Reaction of the inter-
mediate I1 with one of the carbonyl groups of the allophanate
molecule leads to the elimination of the respective carbamate or
alcohol. This is followed by ring closure to isocyanurate.
Scheme 5 Competitive pathways of the reaction network of isocyanu-
rate formation.
Due to the significant progress in the development of more
active and selective catalysts for isocyanurate formation,22 a
quantitative leap concerning catalysts that provide increased
high concentration of allophanate ought to promote pathway activity and improved selectivity to isocyanurate compared to
B. As shown above, isocyanurate formation accelerates once conventional catalysts ought to be feasible. This study concern-
the concentration of allophanate rises to an 8% molar ratio. ing the pathway of isocyanurate formation might provide fresh
More detailed analysis of the kinetic profiles allows for first impetus to generate the necessary new ideas of how to tune
insight into the rate determining steps. The spectra clearly the catalytic cycle.
revealed the prevalence of I1 and allophanate as reaction inter-
mediates, while I2 and I3 were not detected. Thus, in the
anionic route, the reaction of I1 with isocyanate to species I2 Conflicts of interest
and, in the allophanate route, the reaction of I1 with allopha-
There are no conflicts to declare.
nate to intermediate I2 or I3 ought to be rate determining. The
actual ring closure step that provides isocyanurate once inter-
mediate I3 had been obtained ought to be quite fast.
Acknowledgements
PU/PIR foam formulation
Financial support from Covestro Deutschland AG and fruitful
To confirm the catalyst activity in a PU/PIR foam formulation, discussions with Dr Bolko Raffel are gratefully acknowledged.
MDI was reacted with polyol in the presence of KOAc (ratio DC thanks Dr Kai Laemmerhold (Covestro Deutschland AG)
3.5 : 1 : 0.002) and pentane as blowing agent. Analysis of for the support given. We sincerely thank Ms. Ines Bachmann-
samples taken from the core and the surface of the foam (see Remy from the ITMC department at RWTH Aachen University
ESI†) revealed the formation of carbamate and isocyanurate. for performing the in situ NMR experiments.
The isocyanurate content was higher in the sample taken from
the core of the PU/PIR foam. Further, residual isocyanate was
observed in both samples. In the foam formulation, the high Notes and references
temperatures up to 200 °C in the core of the foam will favour
the allophanate route compared to the anionic route (NMR 1 The Polyurethanes Book, ed. D. Randall and S. Lee, Wiley,
characterization of allophanate as intermediate in PU system 2002.

4898 | Polym. Chem., 2018, 9, 4891–4899 This journal is © The Royal Society of Chemistry 2018
Polymer Chemistry Paper

2 L. Nicholas and G. T. Gmitter, J. Cell. Plast., 1965, 1, 85–89. 21 G. Heilig, R. Wiedermann and W. Schmitz, DE 4020255,
3 H. A. Duong, M. J. Cross and J. Louie, Org. Lett., 2004, 6, Bayer A.G., 1992.
4679–4681. 22 A. Al Nabulsi, T. E. Müller, C. Gürtler, H. Vogt, B. Köhler
4 D. K. Hoffman, J. Cell. Plast., 1984, 20, 129–137. and W. Leitner, WO 2016092018, Covestro, 2016.
5 G. W. Ball, G. A. Haggis, R. Hurd and J. F. Wood, J. Cell. 23 M. J. Shaw and W. E. Geiger, Organometallics, 1996, 15, 13–
Plast., 1968, 4, 248–261. 15.
6 M. G. Dekamin, K. Varmira, M. Farahmand, S. Sagheb-Asl 24 A. J. Michel, J. E. Puskas and L. B. Brister, Macromolecules,
and Z. Karimi, Catal. Commun., 2010, 12, 226–230. 2000, 33, 3518–3524.
7 T. Nawata, J. E. Kresta and K. C. Frisch, J. Cell. Plast., 1975, 25 M. A. Thomson, P. J. Melling and A. M. Slepski, Abstr. Pap.
11, 267–278. Am. Chem. Soc., 2001, 221, U316–U316.
8 J. C. Shi, Z. Q. Guo, X. H. Wei, D. S. Liu and M. F. Lappert, 26 S. D. Lipshitz and C. W. Macosko, J. Appl. Polym. Sci., 1977,
Synlett, 2011, 1937–1939. 21, 2029–2039.
9 J. Tang, T. Mohan and J. G. Verkade, J. Org. Chem., 1994, 27 E. Broyer, C. W. Macosko, F. E. Critchfield and L. F. Lawler,
59, 4931–4938. Polym. Eng. Sci., 1978, 18, 382–387.
10 S. J. Peters, J. R. Klen and N. C. Smart, Org. Lett., 2008, 10, 28 G. H. Lorimer, Trends Biochem. Sci., 1983, 8, 65–68.
4521–4524. 29 D. Seo and J. R. Youn, Polymer, 2005, 46, 6482–6493.
11 H. E. Reymore, P. S. Carleton, R. A. Kolakowski and 30 M. Thirumal, D. Khastgir, N. K. Singha, B. S. Manjunath
A. A. R. Sayigh, J. Cell. Plast., 1975, 11, 328–344. and Y. P. Naik, J. Appl. Polym. Sci., 2008, 108, 1810–
12 H. Ulrich, B. Tucker and A. A. R. Sayigh, J. Org. Chem., 1817.
1967, 32, 3938–3941. 31 S. Pinchas and D. Ben-Ishai, J. Am. Chem. Soc., 1957, 79,
13 K. Schwetlick and R. Noack, J. Chem. Soc., Perkin Trans. 2, 4099–4104.
1995, 395–402. 32 R. R. Romero, R. A. Grisby, E. L. Rister, J. K. Pratt and
14 A. Farkas and G. A. Mills, in Advances in Catalysis, D. Ridgway, J. Cell. Plast., 2005, 41, 339–359.
ELSEVIER, 1962, vol. 13, ch. 6, pp. 393–446. 33 M. Yamashita and J. B. Fenn, J. Phys. Chem., 1984, 88,
15 D. P. N. Satchell and R. S. Satchell, Chem. Soc. Rev., 1975, 4, 4451–4459.
231. 34 M. Yamashita and J. B. Fenn, J. Phys. Chem., 1984, 88,
16 K. C. Frisch and L. P. Rumao, J. Macromol. Sci., Rev. 4671–4675.
Macromol. Chem., 1970, 5, 103–149. 35 J. F. Banks, S. Shen, C. M. Whitehouse and J. B. Fenn, Anal.
17 K. Schwetlick, R. Noack and F. Stebner, J. Chem. Soc., Perkin Chem., 1994, 66, 406–414.
Trans. 2, 1994, 599–608. 36 S. Friso, S. W. Choi, G. G. Dolnikowski and J. Selhub, Anal.
18 The Chemistry of Cyanates and Their Thio Derivatives, ed. R. Chem., 2002, 74, 4526–4531.
Richter and H. Ulrich, John Wiley & Sons, Chichester, 1977. 37 M. Moini, B. L. Jones, R. M. Rogers and L. Jiang, J. Am. Soc.
19 K. Dusek, M. Spirkova and I. Havlicek, Macromolecules, Mass Spectrom., 1998, 9, 977–980.
1990, 23, 1774–1781. 38 K. A. Chaffin, A. J. Buckalew, J. L. Schley, X. Chen, M. Jolly,
20 M. Funatsu, S. Yasuda and T. Hiraizumi, JP 50072998, J. A. Alkatout, J. P. Miller, D. F. Untereker, M. A. Hillmyer
US3953384, Kao Corp., 1975. and F. S. Bates, Macromolecules, 2012, 45, 9110–9120.

This journal is © The Royal Society of Chemistry 2018 Polym. Chem., 2018, 9, 4891–4899 | 4899

View publication stats

You might also like