Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Macromol. Chem. Phys.

198,117-134 (1997) 117

Infrared temperature studies of a simple polyurea


Michael M. Coleman*, Maria Sobkowiak, George J. Pehlert, Paul C. Painter
Department of Materials Science and Engineering, The Pennsylvania State University,
University Park, PA 16802, U. S . A.

Tahir Iqbal
E. I. du Pont de Nemours & Co., Inc., Wilmington, DE 19880, U. S. A.
(Received: March 21, 1996; revised manuscript of July 5, 1996)

SUMMARY
Infrared temperature studies of a polyurea formed from 2-methylpentane-1,5-diamine
and 4,4’-methylenebis(pheny1 isocyanate) are presented, together with those of two
model diureas, 2-methylpentane-1,5-diphenylureaand ethylene diphenylurea. The main
thrust of the work concerns the interpretation of the infrared changes that are observed in
ordered and disordered hydrogen bonded ureas.

Introduction
We have previously reported detailed infrared spectroscopic studies of amorphous
and semi-crystalline polyamides ’), a simple (non-segmented) polyurethane2) and
polyether blends with amorphous polyamides and polyurethanes3). The focal point
of this work was centered upon hydrogen bonding and the interpretation of the infra-
red spectra of polyamides and polyurethanes as a function of temperature, princi-
pally in the N-H and C=O (Amide 1) stretching regions. At a particular tempera-
ture, in the amorphous state above the glass transition temperature (Tg),polyamides
and polyurethanes form an equilibrium distribution of chain-like linear hydrogen
bonded structures of varying length (denoted B, where h = 0, 1, 2, 3 , etc.) as illu-
strated below in Scheme 1 (which depicts a typical B3trimer).
The system is dynamic, with hydrogen bonds rapidly forming and breaking at the
urgings of thermal motion. Infrared spectroscopy captures a representative structure
that reflects an average equilibrium distribution of the hydrogen bonded “chains”. In
the N-H stretching region ( 3 100-3700 cm-l) we observe a very broad band in the
spectra of polyamides and polyurethanes, characteristic of the wide distribution of
hydrogen bonds (of different distances and geometries) that would be expected in
disordered or “amorphous” hydrogen bonded structures‘. ’). “Free” (non-hydrogen
bonded) N-H groups that are present at the end of the hydrogen bonded “chains”
give rise to a relatively narrow N-H stretching band observed at the high wavenum-
ber edge of the distribution. However, the absorption coefficients of such “free”
N-H bands are known to be small compared to their hydrogen bonded counterparts
making them difficult to detect’.’). In the carbonyl (or Amide 1) stretching region,
two corresponding bands are observed that are also attributed to disordered hydro-
gen bonded and “free” carbonyl groups, with the former, in this case, only being
somewhat broader than the latter.
0 1997, Huthig & Wepf Verlag, Zug CCC 1022-1352/97/$10.00
118 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P. C. Painter, T. Iqbal

Scheme I :

“free”C=O

Polyamide Polyurethane Polyurea

The effect of order (“crystallinity”) on the N-H and C=O stretching regions of
the infrared spectra of polyamides and polyurethanes is profound and distinct for the
two separate regions. First, in contrast to the very broad hydrogen bonded N-H
stretching band observed in the spectra of amorphous (or disordered) polyamides
and polyurethanes, the spectra of the analogous semi-crystalline (ordered) polymers
is dominated by a much narrower overall band envelope characteristic of a corre-
spondingly narrower distribution of hydrogen bonded distances and geometries.
However, if the band envelope is examined carefully a contribution from disordered
(amorphous) hydrogen bonded N-H groups is also generally observed, as evi-
denced by a broad underlying band superimposed on the relatively sharp ordered
hydrogen bonded N-H band. Hence, the band envelope in the N-H stretching
region of pol yamides and polyurethanes reflects the overall distribution of hydrogen
bonded N-H groups. In other words, there is no distinction between an >N-H- - -
O=C< hydrogen bond of a particular strength and geometry, according to whether it
is present in an ordered or disordered hydrogen bonded domain.
This is not the case in the carbonyl stretching region. Here, in the spectrum of
semi-crystalline polyamides and polyurethanes, there are two distinct bands corre-
sponding to ordered and disordered hydrogen bonded carbonyl groups. In simple
terms, dipole/dipole coupling, resulting from the regular alignment of the amide or
urethane carbonyl groups arranged in a two or three dimensional lattice, produces a
shift of the infrared active carbonyl stretching vibration (Amide 1) in ordered hydro-
gen bonded structures to lower frequency. Accordingly, the carbonyl stretching
region of simple polyamides and polyurethanes is informationally rich and it is pos-
sible to quantitatively determine the fraction of amide or urethane carbonyl groups
Infrared temperature studies of a simple polyurea 119

in ordered and disordered hydrogen bonded domains, together with those that are
“free” (non- hydrogen bonded)‘. 2,
In this current work we turn our attention to a simple (non-segmented) polyurea.
It is interesting from both structural and spectroscopic viewpoints that urea moieties
contain one hydrogen bonding acceptor (the C=O group) and two donors (the two
N-H groups). Thus, in contrast to the linear 1: 1 hydrogen bonds that are formed in
amides and urethanes, the self-association of urea groups involves chain-like non-
linear 1 :2 bifurcated hydrogen bonds formed between the single C=O and both
N-H groups (see Scheme I )5). However, unlike polyamides and polyurethanes,
polyureas have not been subjected to intense scrutiny. Simple polyureas that are ana-
logous to the aliphatic or aromatic nylons tend to be insoluble, semi-crystalline
materials with very high melting points (T,). For example, the T, of polyhexa-
methyleneurea is =295”C, while the analogous polyamide (nylon 6 ) has a T, of
21 5 O C 4 ) . Nevertheless, urea (and urethane) groups are important constituents of the
more complex segmented polyether multiblock copolymers (SPUU), and there have
been a number of infrared spectroscopic studies on these complex But
before attempting to unravel the myriad of possible different hydrogen bonding
interactions in these materials we believe it is necessary to first fully understand
those that occur in simple polyureas.

Experimental part
N,N-Dimethylformamide (DMF), toluene, dimethyl sulfoxide (DMSO), N,N-dimethyl-
acetamide (DMAC), 1,1,1,3,3,3-hexafluoro-2-propanol(HFIP), ethanol, phenyl isocya-
nate (PI), ethylene diamine (EDA) and 4,4’-methylenebis(pheny1isocyanate) (MDI) were
purchased from Aldrich Chemical Co. Inc. 2-Methylpentane-l,5-diamine(MPDA) was
supplied by DuPont.
Ehylene diphenylurea (EDPU) was prepared in the following manner. A solution of
24.990 g (210.0 mmol) PI in 125 mL anhydrous DMAC was cooled in an ice bath under
nitrogen. To this stirred solution was added dropwise a solution of 6.000 g (100.0 mmol)
ethylene diamine in 10 mL of anhydrous DMAC. The temperature of the reaction mix-
ture was kept between 5- 10°C during the addition of the diamine solution. The white
solid product was filtered and washed 3 times with 50 mL ethanol under suction filtra-
tion. It was then dried under vacuum to a constant weight, 20.87 g (70% yield). A similar
procedure was used to prepare 2-methylpentane-l,5-diphenylurea(MPDPU). A solution
of 12.495 g (105.0 mmol) PI in 75 mL toluene was cooled in an ice bath under nitrogen.
To this stirred solution was added dropwise a solution of 5.800 g (50.0 mmol) of MPDA
in 15 mL toluene. The temperature of the reaction mixture was kept between 5 - 10°C
during the addition of the diamine. The white solid product was filtered, washed 3 times
with 25 mL toluene and air dried, 16.28 g (92% yield).
The simple polyurea formed from the reaction of MDI and MPDA, which we denote
PU, was synthesized as follows. A solution of 2.500 g (10.0 mmol) MDI in 50 mL anhy-
drous DMAC contained in a 250 mL 3-neck round bottom flask was cooled in an ice
bath to between 5 1 0 ° C under nitrogen. To this stirred solution was dropwise added a
solution of 1.195 g (10.3 mmol) MPDA in10 mL DMAC. The rate of the diamine addi-
tion was controlled to keep the reaction mixture below 10°C. The contents of the flask
were then slowly poured into 500 mL water with vigorous stirring. The solution was
further stirred for three hours and filtered. The precipitate was first washed with water
120 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P. C. Painter, T. lqbal

and then two times with 50 mL ethanol. Overnight drying under vacuum afforded
3.184 g (87% yield) of a white solid product.
Infrared spectroscopic measurements were recorded on a Digilab model FTS60 Fourier
transform infrared (FTIR) spectrometer at a resolution of 2 cm-’. Films of the model
ureas MPDPU and EDPU, the homopolyurea PU and the blends of PU with PEO for
transmission infrared analysis were cast on KBr windows principally from DMF solu-
tions. Details of the solvent used, solution concentration and solvent evaporation tem-
perature and rate are mentioned at appropriate positions in the text. Transmission spectra
recorded at elevated temperatures were obtained using a heating cell horizontally
mounted inside the sample chamber. Temperature was regulated by a Micristar 828D
digital process controller, which has a reported accuracy of kO.1 “C. Special attention was
paid to ensure that all FTIR samples were sufficiently thin to be within the absorption
range where the Beer-Lambert law is obeyed. Thermal analysis was conducted on a
Seiko Instruments differential scanning calorimeter (DSC-220CU) coupled to a SSC
5300 computerized data station. A heating rate of 2O0C/min was used and the glass tran-
sition temperatures recorded as the midpoint of the heat capacity change.

Results and discussion

Preamble
We chose to study the simple (non-segmented) polyurea formed from the reaction
of 2-methylpentane- 1,Sdiamine (MPDA) and 4,4‘-methylene bis(pheny1 isocyanate)
(MDI), which we denote PU, for reasons that in retrospect now appear rather naive.

(PU)
In common with our previous studies on amorphous polyamides’as’b)and poly-
urethanes3), we believed it beneficial to commence infrared temperature experiments
on a sample of a soluble amorphous polyurea, one which should not order (“crystal-
lize”) and complicate the initial interpretation of the infrared spectra. Unlike a sym-
metrical diamine, such as ethylene diamine (EDA), which produces a highly crystal-
line polyurea when reacted with MDI, we reasoned that the asymmetric MPDA
should yield an amorphous polyurea, because of the disorder caused by the irregular
placement of MPDA segments into the polymer chain, in terms of both stereo- and
sequence isomerism’). As we will see, things are not that simple. The PU polymer is
readily soluble in solvents such as DMF, DMAC, DMSO, HFIP and the like and
does form amorphous films from these solvents under “normal” conditions. How-
ever, order in the PU sample can be induced by varying the rate of solvent evapora-
tion and the temperature at which it occurs. Moreover, in dilute solutions of PU in
DMF we observe the interesting phenomenon of the slow formation of gels at ambi-
ent temperature and this is related to the development of order. But before we exam-
ine the results obtained from the PU polymer, it is useful to study the infrared spec-
tra of two relevant low molecular weight model diureas.
Infrared temperature studies of a simple polyurea 121

Room temperature infrared studies of two model compounds


The chemical structures of two relevant low molecular weight materials, 2-
methylpentane- 1,5-diphenylurea (MPDPU), a fine model for the above PU polyurea,
and ethylene diphenylurea (EDPU), included because it should be highly ordered,
are shown below:

H H H H

II
0 (MPDPU) 0

H H 0

H H
(EDPU)

Fig. 1 compares the room temperature infrared spectra from 500-4000 cm-’ of
MPDPU and EDPU (denoted A and B, respectively), cast as thin films from a 2%
DMF solution onto KBr windows. While there are many features, spread across the
entire spectral range, that reflect differences in the chemistry and physical state of
the two model ureas, we focus our attention in this study on the N-H (3 100-3600
cm-I) and urea C=O (1 600- 1 750 cm-’) stretching regions. Scale expanded spectra
in these specific regions are shown in Fig. 2.
Let us first consider the spectrum of MPDPU. The very broad band seen in the
spectrum of MPDPU centered at =3 340 cm-’ (Fig. 2A) is characteristic of a broad
distribution of hydrogen bonds of different distances and geometries and is typical
of disordered or “amorphous” hydrogen bonded structures such as those observed in
pol yamides and polyurethanes lW3).There is no evidence of any “free” (non-hydrogen
bonded) N-H groups that would give rise to a band at ~ 3 4 5 0 cm-I. However, as
noted in the introduction, such “free” N-H bands are difficult to detect because
their absorption coefficients are known to be small compared to their hydrogen
bonded counterparts’.’). The two small bands between 3 100 and 3200 cm-‘ are pre-
sumably overtone or combination bands. In the carbonyl stretching region, which is
analogous to the Amide 1 band of polyamides and polypeptides, the ~ 1 6 5 cm-’ 0
mode is assigned to disordered hydrogen bonded carbonyl groups and the shoulder
at ~ 1 6 9 cm-I
0 to “free” (non-hydrogen bonded) carbonyl groups. Note that these
bands occur at lower frequencies and are conveniently separated from those attribu-
ted to the analogous disordered hydrogen bonded and “free” urethane carbonyl
groups2). To summarize, the infrared spectra recorded in the N-H and C=O stetch-
ing regions establishes that the thin film of MPDPU, as prepared, is essentially an
amorphous material at room temperature.
The infrared spectrum of EDPU is quite different. First the relatively narrow band
envelope seen in the N-H stretching region (Fig. 2B) is characteristic of a corre-
spondingly narrow distribution of hydrogen bond lengths, typical of that observed
122 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P. C. Painter, T. Iqbal

m 3500 3000 2500 2000 I500 1000 503


Wavenumber in crn-’
Fig. 1. Infrared spectra of cast films of (A) MPDPU and (B) EDPU recorded at room
temperature in the region 500-4000 cm-’

, 1
3600 3400 3203 if50
1750 1700 1650 1600
16bO
Wavenumber in cm-’ Wavenumber in crn-’
Fig. 2. Scale expanded infrared spectra of (A) MPDPU and (B) EDPU recorded at
room temperature on the left: N-H stretching region (3 100-3600 cm-’) and right:
C-0 stretching region (1 600- 1750 cm-’)

for ordered (“crystalline”) hydrogen bonded structures172).Furthermore, the doublet


observed at -3 320/3 340 cm-’ is most probably attributed to the in-phase and out-
of-phase N-H stretching vibrations of the two N-H groups of the highly ordered
urea groups. There is also some evidence of a contribution from disordered (amorph-
ous) hydrogen bonded N-H groups present as a broad underlying band. However,
we again see no evidence of any “free” (non-hydrogen bonded) N-H groups.
Infrared temperature studies of a simple polyurea 123

In the carbonyl stretching region the spectrum is dominated by the band at


-1 630 cm-’, which, in common with the trends observed in ordered polyamide and
polyurethanes’, 2), is assigned to ordered (“crystalline”) hydrogen bonded carbonyl
groups. There is little evidence of a contribution at 1650 and 1690 cm-’ (disordered
hydrogen bonded and “free” carbonyl groups, respectively). This is all consistent
and indicates that the urea groups in the thin film of EDPU, as prepared, are essen-
tially ordered or “crystalline” at room temperature.
A summary of the relevant infrared assignments for the urea group is presented in
Scheme 2 below.

Scheme 2: Characteristic infrared bands for urea groups

uC’O= 1690 cm‘’

H H

“Free”(Non Hydrogen Bonded)


/‘,
I1
/
(Ordered-sharp)
=3340cm-’
Urea Group ‘ yH. , .Y (Disordered-broad)

,* . .
Urea/Urea Hydrogen Bond

The most important conclusion of these initial room temperature infrared studies
on the model ureas is that in common with amides and urethanes, we can clearly
distinguish between ordered and disordered hydrogen bonded and “free” urea
groups.

Infrared temperature studies of the model MPDPU


For reasons expressed in the above preamble, we expected that infrared tempera-
ture studies of both the MPDPU and EPU model diureas would be rather mundane.
Infrared spectra of the latter recorded as a function of temperature were expected to
remain characteristic of an ordered (“crystalline”) material up to the melting point
or onset of significant degradation. This is essentially correct, at least up to =20O0C.
As we increase the temperature a slight broadening of the hydrogen bonded N-H
stretching band is observed and the in-phase and out-of-phase N-H stretching
vibrations are no longer resolved (see Fig. 2). Concurrently, the ordered urea carbo-
nyl band at ~ 1 6 3 cm-’
0 shifts by a few wavenumbers to a higher frequency and
also broadens perceptibly. This is what we would expect to see in the spectrum of a
“crystalline” material resulting from an increase in the thermal motion as the tem-
124 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P.C. Painter, T. Iqbal

perature is raised, so we will not dwell further on the infrared spectra of EPU. On
the other hand, infrared spectra of MPDPU were expected to remain characteristic
of an amorphous material and simply show a change in the ratio of the “free” to
hydrogen bonded carbonyl groups as a function of increasing temperature, together
with some typical shifting and broadening of the bands, especially those associated
with hydrogen bonded groups. As we will see below, however, infrared temperature
studies of MPDPU were far more interesting and complex.
Figs. 3 and 4 show scale expanded infrared spectra of a sample of MPDPU cast as
a thin film from an -2% DMF solution onto a KBr window, recorded in the N-H
stretching (3 100-3600 cm-’) and carbonyl stretching regions (1 600-1 750 cm-’),
respectively, recorded while heating from ambient to 20°C in 10°C increments.
From room temperature to 80 “C the spectra remain consistent with that of an amor-
phous material, with a broad N-H stretching band and two bands at -1650 and
1690 cm-’, assigned to disordered hydrogen bonded and “free” urea carbonyl
groups, respectively. At 90°C, however, there is a discernible narrowing of the
N-H band (Fig. 3) and evidence of a shoulder at ~ 1 6 3 0cm-’ in the carbonyl
stretching region (Fig. 4), which suggests the urea groups in MPDPU, which were
disordered (amorphous) as prepared, start to order (“crystallize”). In the spectrum
recorded at IOO’C, a further narrowing of the N-H band is obvious and the carbo-

Fig. 3. Scale expanded


infrared spectra of a film
of MPDPU recorded as a
??% function of increasing tem-
perature from 80 to 200 “C
80°C fi in the NH stretching region
(3 100-3600 cm-’)
I . , I
3600 3500 3400 3300 3200 3100
Wavenumber in cm-’
Infrared temperature studies of a simple polyurea 125

Fig. 4. Scale expanded


infrared spectra of a film
of MPDPU recorded as a
function of increasing tem-
perature from 80 to 200 “C
in the C=O stretching
region (1 600- 1 750 cm-’)
1750 1700 1650 1600
Wavenurnber in crn-’

nyl band at -1630 cm-I is now dominant implying that the urea groups are mainly
in ordered hydrogen bonded structures. However, there is still an obvious contribu-
tion from the bands at -1 650 and 1690 cm-’, indicating a significant fraction of the
urea groups are still in disordered structures. Within the temperature range of 100 to
160“C, the urea carbonyl groups appear to be essentially ordered (“crystalline”), as
there is no obvious sign of any contributions from the disordered hydrogen bonded
or “free” urea carbonyl bands at -1650 and 1690 cm-’ (Figs. 3 and 4).At 170°C
and up to 200°C the MPDPU sample revert to a disordered (amorphous) material.
The N-H band envelope broadens enormously (Fig. 3) and resembles that observed
at 80°C and below (albeit shifted somewhat to higher frequency). Concurrently, we
also see the return of the disordered hydrogen bonded and “free” urea carbonyl
bands at ~ 1 6 5 and
0 1690 cm-l (Fig. 4) and the absence of any significant contribu-
tion from the ordered band at -1630 cm-l.
To summarize to this point, the MPDPU sample, which was amorphous as pre-
pared, orders or “crystallizes” at approximately 90 “C. Given the irregular aliphatic
sequence between the two urea groups of MPDPU, this was unexpected (at least to
the authors). Above approximately 160 “C the ordered (“crystalline”) urea reverts
back to a disordered (“amorphous”) material and, in effect, the ordered urea struc-
ture “melts”.
126 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, I? C. Painter, T. Iqbal

It is reasonable to assume that kinetic factors are important and the degree of
order attained at a given temperature will also depend how long the sample is held
at that temperature. This was tested by heating a film of MPDPU to 90°C and
recording the infrared spectrum after 15, 30,45 and 60 min. The relative absorbance
of the carbonyl band attributed to ordered hydrogen bonded urea structures was
indeed observed to slowly increase indicating the development of ordered hydrogen
bonded urea groups. In a separate experiment MPDPU was heated quickly from
25°C to 150°C and spectra were recorded immediately, after 15 min at 150°C and
finally at 25°C after cooling to room temperature. The spectra, presented in Fig. 5,
show unambiguously that in the initial spectra recorded at 25 “C, and immediately
upon attaining a temperature of 150°C the urea groups are disordered (“amor-

150°C - 15 min

150°C - Initial

25Ta ,/ <
3& 3i00 3d00 3300 3200 3100
Wavenumber in cm-’

Fig. 5. Scale expanded


infrared spectra of a film
of MPDPU recorded at
25 “C, immediately after
heating to 150°C, after
15 min at 150°C and
150°C - 15 min finally cooling back to
25 “C. Top: the N-H
150°C- Initial stretching region
L____
(3 100-3600 cm-I);
bottom: C=O stretching
region (1 600- 1 750 cm-’)
1750 1700 1650
Wavenumber in crn-’

phous”), but that after 15 min at 150°C and upon cooling to room temperature a
fraction of the urea groups are transformed to ordered (“crystalline”) structures.
Fig. 6 shows the results of a similar study performed at 160”C, close to the “melt-
ing“ point of ordered urea structures. The sample was heated quickly from 25 “C to
Infrared temperature studies of a simple polyurea 127

2 , . , ,

4
3600 3500 3400 3300 3200 3100
Wavenumber in crn-'

Fig. 6. Scale expanded


infrared spectra of a film
of MPDPU recorded at
25 "C, immediately after
heating to 160"C and after
15,30,45 and 60 min at
160°C. Top: the N-H
stretching region (3100 -
3 600 cm-'); bottom: C=O 160°C - Initia
stretching region (1 600-
1750 cm-') 25OC
I I
1750 1700 1650 1600
Wavenumber in crn-'

160°C and spectra were recorded immediately and after annealing for 15, 30, 45
and 60 min at 160°C. In essence, these spectra show that in the initial spectra
recorded at 25°C the urea groups are essentially disordered, but that as soon as a
temperature of 160°C is attained a large fraction of the urea groups become ordered.
However, with increasing time at 160°C the fraction of ordered urea groups
decreases. This suggests that 160°C is just above the ordeddisorder transition and
the ordered structures formed while heating to 160°C are now disordering.
Returning to the primary temperature cycle, it is important to recognize that the
phenomenon described in the heating study from room temperature to 200 "C, (i. e.
disordered to ordered to disordered urea groups), was not reversible. Under the
experimental conditions employed, the urea carbonyl groups do not order ("recrys-
tallize") when the sample is slowly cooled from 200°C back to room temperature,
as evidenced by the spectra shown in Figs. 7 and 8. The N-H stretching envelope
remains very broad and only the bands at ~ 1 6 5 and0 1690 cm-' are observed in the
carbonyl stretching region of the infrared spectra.
An attempt was made to see if it was possible to develop ordered hydrogen
bonded urea structures from the "melt" (above 170°C) by annealing at a temperature
128 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P. C . Painter, T. Iqbal

L,
3600 3500 3400 3300 3200 3100 17jo 1iw It50 16bo

Wavenumber in cm-’ Wavenumber in cm”


Fig. 7. Fig. 8.

Fig. 7. Scale expanded infrared spectra of a film of MPDPU recorded as a function of


decreasing temperature from 200 to 100°C in the N-H stretching region (3100 -3600
cm-’1
Fig. 8. Scale expanded infrared spectra of a film of MPDPU recorded as a function of
decreasing temperature from 200 to 100°C in the C=O stretching region (1 600- 1750
cm-’)

below the orderldisorder transition. A new sample of MPDPU was heated rapidly in
the spectrometer to 170°C and the spectrum recorded immediately and after a period
of 15 min. Again, consistent with the results presented above, there was no evidence
of ordered (“crystalline”) hydrogen bonded ureas. Next the sample was cooled to
140°C, a temperature at which ordered hydrogen bonded ureas were observed to be
present in the heating cycle (Fig. 4), and the sample annealed at 140°C for a time
period of up to 60 min. Once more, there was no evidence for any ordering of the
urea groups. In a similar vein, a film cast from MPDPU was heated rapidly to
170“C, cooled to room temperature and a reheating study performed, similar to that
employed to obtain the spectra shown in Figs. 3 and 4, to see if order could be
induced in a sample that has been exposed to such a temperature. It could not.
Finally, a film cast from MPDPU was heated rapidly to 170°C cooled to room tem-
perature, redissolved in DMF, recast, and then a heating study performed. Again, no
evidence of ordering was found.
Fig. 9 shows DSC thermograms obtained from a sample of MPDPU subjected to
the following thermal treatment. The sample was first heated from room temperature
Infrared temperature studies of a simple polyurea 129

Fig. 9. DSC curves of


MPDPU. First run: sample
heated from -80 "C to
250°C at 20"C/min and
then cooled to ambient
temperature. Second run:
sample reheated from
50 "C to 250 "C at
20 Wmin 1
-50
.. .. 1

0
....
50
I . . . I

100
Temp. in "C
I

150
172C

. .. .. ... . . . . . .i
200

to 100°C (to reduce "settling" problems in the DSC pan), cooled to -80°C and then
heated at 20"C/min to 250°C (denoted first run). The Tg (17°C) and T, (onset
-16OoC, maximum 172°C) of the MPDPU sample are clear and unambiguous. The
sample was then cooled to 50°C and the heating cycle repeated (second run) at
20"C/min to 250 "C. This time no melting transition was detected.
The above is strong evidence for the occurrence of chemical changes to MPDPU
in the melt (at temperatures >160°C). Given that there are few obvious changes in
the infrared spectra of amorphous samples of MPDPU before and after thermal treat-
ment above 160"C, (there is, however, some evidence for an unassigned additional
5 - see Fig. 8), we believe that
band in the carbonyl stretching region at ~ 1 7 0 cm-'
the most probable and prevalent chemical changes occurring in the MPDPU melt
are transurea reactions of the type shown below in Scheme 3. Under these circum-
stances, the original sample of monodisperse MPDPU diureas are transformed into a
distribution of urea oligomers (i.e. mono-, di-, triureas, etc.) which would tend to
thwart the development of order ("crystallization"). Note that in MPDPU diureas all
the urea groups are of the aromatic/aliphatic type, where each is attached to one phe-
nyl and one methylene group. In the early stages of transurea reactions diureas are
transformed to aromatic/aromatic (monoureas) and a 2 : 1 mixture of aromatic/ali-
phatic and aliphatic/aliphatic (triureas). The "free" carbonyl stretching frequency for
aromatic/aromatic urea groups would be expected to occur at a higher frequency
than that of the corresponding aromatic/aliphatic urea which is observed at ~ 1 6 9 0
cm-' (and at an even higher frequency than that of the corresponding aliphatidali-
phatic urea). Thus the detection of the ~ 1 7 0 5cm-' band mentioned above in
MPDPU samples that have been exposed to temperatures above 160°C is consistent
with the formation of monoureas. We have reported on similar infrared studies of
analogous trans reactions in polycarbonates').
130 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P. C. Painter, T. Iqbal

Scheme 3 :

N-C-N-R-N-C-N
II II
0 0

+
- H H H H H H -

Infrared studies of the polyurea PU


As we intimated in the preamble to this section of the paper, we originally thought
that a simple PU polyurea synthesized from MPDA and MDI would most probably
be incapable of crystallizing to any measurable extent because of the structural irre-
gularity present in the polymer chain caused by the incorporation of both stereo- and
sequence isomers of the MPDA during polymerization. Of course, this was before
we had performed the infrared spectroscopic experiments on the MPDPU model pre-
sented above, which if nothing else suggests that our initial elementary hypothesis
of an amorphous uncrystallizable PU may be erroneous. Nevertheless, the first set of
results obtained from films of PU cast from 1% DMF solutions did point to an essen-
tially amorphous polyurea.
The spectrum in Fig. 10, which is characteristic of an amorphous polyurea, (with
a relatively broad N-H stretching envelope and two carbonyl stretching bands at
-1 650 and 1690 cm-’, assigned to disordered hydrogen bonded and “free” urea car-
bony1 groups, respectively, together with no obvious contribution from ordered
hydrogen bonded urea carbonyl groups at ~ 1 6 3 cm-’),
0 was obtained from a PU
film sample prepared by evaporating the solvent relatively quickly from a 1% solu-
tion in DMF. This PU film was cast from solution at a temperature of 80°C onto a
KJ3r plate maintained at this temperature. After evaporation of the majority of the
solvent, the sample was placed into a vacuum oven at 80°C for 4 h to completely
remove the remaining solvent (in practice, until the solvent was not detected by
infrared spectroscopy). Identical spectra were also obtained from films prepared at
Infrared temperature studies of a simple polyurea 131

room temperature in an open fume hood, followed by vacuum desiccation at room


temperature for 4 d. Moreover, films prepared in the same manner from DMSO or
HFIP yielded similar spectra. Heating these films to 200°C produced little differ-
ence in the spectrum in the primary N-H and C = O stretchings regions (apart from
the gradual frequency shifts typically observed with temperature), as shown in the
inserts of Fig. 10. Unlike the analogous spectra of the MPDPU model compound
(Figs. 3 and 4), there is no evidence for the development of order as a function of
temperature. This is hardly surprising since the Tg of PU from thermal analysis is
-135OC (see later). As prepared the PU film sample is amorphous and if crystal-
lization of PU was possible it would be necessary to anneal above the Tg in the
“melt” at temperatures in excess of 200 “C, where presumably chemical (transurea)
reactions would also be expected to be occur.

0.25

a,
0

es
8

0.05
4000 3iw 3doo 2i00 2doo 1500 loo0 500

Wavenumber in cm-’
Fig. 10. Infrared spectrum recorded at room temperature in the region 500-4000 cm-l
of a film of PU formed by rapid evaporation from a 1% solution in DMF. Inserts: scale
expanded spectra recorded at 40, 80, 120, 160 and 200°C in the N-H (3 100-3600
cm-’) and C=O (1 600- 1750 cm-l) stretching regions

In previous infrared studies of polymer blends3“.l o ) we have observed the develop-


ment of crystallinity in polycarbonate and aromatic/aliphatic polyamides that was
induced by a solvent (of low or high molecular weight). This is a well known phe-
nomenon and in essence the presence of the solvent decreases the overall Tg of the
system, depresses the crystalline melting point, Tm, and brings the crystalline rate
curve down to the point where the development of crystallinity becomes plausible at
room temperature. Accordingly, we postulated that if PU was indeed capable of
crystallization (as implied by the MPDPU studies), it might be possible to induce
the same by manipulation of the solution concentration, evaporation rate, etc.
Fig. 11 shows a spectrum of PU that is characteristic of an ordered (“crystalline”)
polyurea (with a relatively narrow N-H stretching envelope and three carbonyl
132 M. M. Coleman, M. Sobkowiak, G. J. Pehlert, P. C. Painter, T. Iqbal

0.85-

5
a,

h
0.60.


2
(I)

0.35-

&
0.10
35ii 3000 2500 zoo0 1500 1000 500
Wavenumber in cm-’

Fig. 11. Infrared spectrum recorded at room temperature in the region 500-4000 cm-’
of a film of PU formed by the very slow evaporation from a 1 % solution in DMF. Inserts:
scale expanded spectra recorded at 40,80, 120, 160 and 200°C in the N-H ( 3 100-3 600
cm-’) and C = O ( I 600- 1 750 cm-’) stretching regions

stretching bands at -1 630 (dominant), 1650 and 1690 cm-’ (shoulders), attributed
to ordered hydrogen bonded, disordered hydrogen bonded and “free” urea carbonyl
groups, respectively) that was prepared by the very slow evaporation of DMF. Spe-
cifically, the solution of PU in DMF solvent was cast at room temperature onto a
KBr window that was placed in a small aluminium pan containing a small pool of
pure DMF and then loosely covered with a larger petri dish. Evaporation of DMF
from the sample was exceedingly slow and extended over a period of days. Heating
the PU film prepared in this manner to 200 “C again produced little difference in the
spectrum in the primary N-H and C=O stretching regions, as illustrated in the
inserts of Fig. 11. Again, this is not surprising since the PU film sample after
removal of the solvent is now “crystalline” with a Tg of =135OC and a T,,, of
=26OoC (see below).
Concurrent with the studies mentioned above, we observed that PU solutions in
DMF gradually form gels as a function of time, concentration and temperature. At a
concentration of ~ 0 . 4 % PU in DMF, PU dissolves readily and forms a clear liquid
(single phase) from which films can be easily prepared. However, after a few days at
ambient temperature there is the obvious presence of gel particles, mostly clinging
to the sides of the sealed container. The presence of gel is accentuated by increasing
the concentration and/or reducing temperature. For example, a 1% solution of PU in
DMF initially forms a clear solution, but within a few hours forms a space filling
gel. An analogous 1 % solution in HFIP does not appear to gel at room temperature,
Infrared temperature studies of a simple polyurea 133

Fig. 12. Comparison of


the infrared spectrum in
the N-H stretching region
(3 100-3600 cm-') of a
sample of the PU gel
formed in a 1% solution of
DMF recorded at room
temperature with the spec-
tra of PU films formed by
the slow and rapid eva-
PU gel ("smeared)
poration from DMF solu- from I % soh in DMF
tion
3600 3500 3400 3300 3 m 3100

Wavenumber in cm-'

but will do so if the sample is replaced in a refrigerator at = -20°C. What are these
gels and what is the driving force for their formation?
Fig. 12 compares the N-H stretching region of the infrared spectra obtained from
three different PU samples. Those prepared by the rapid and slow evaporation of
DMF from clear 1% solutions give spectra similar to those described above that are
characteristic of disordered (amorphous) and ordered ("crystalline") polyureas,
respectively. The third spectrum is that obtained from the gel formed in the original
DMF solution after storing overnight at room temperature. Some of this gel was
"smeared" onto a KBr window and the spectrum recorded. Fortunately, DMF does
not absorb significantly in the 3 100-3 600 cm-' region of the infrared spectrum (in

Fig. 13. DSC curve of a


sample of the dried PU gel
formed in a 1% solution of
DMF. The sample heated
from room temperature to
300 "C at 20 "C/min
134 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P. C. Painter, T. Iqbal

contrast to the 1600- 1750 cm-’ region, where it does) and it was not necessary to
completely evaporate the solvent. From the breadth of the N-H stretching envelope
in the spectrum of the smeared gel we can determine unambiguously that the urea
groups in the gel are predominantly ordered. This was confirmed by thermal analysis
of a dried sample of the gel which is presented in Fig. 13 and shows a clear T, of
258°C. Thus we conclude the gel is produced by the relatively slow formation of
ordered (microcrystalline) urea domains and these are presumably linked together
by solvent swollen amorphous chains. These ordered domains are the junction points
necessary to provide the connectivity and physical integrity of the gel.

Polyuredpolyether blends
Finally, we will briefly consider some infrared results obtained from films of PU
blends with poly(ethy1ene oxide) (PEO), as they have a direct bearing on spectral
interpretation. Analogous studies of polyether blends with amorphous polyamides
and polyurethanes have previously been reported3). As mentioned before, both poly-
amides and polyurethanes strongly self-associate (see Scheme I), but in mixtures
with polyethers the oxygen atoms also compete for the N-H groups to form
N-H- - -0 hydrogen bonds (termed interassociation). The phase behavior of hydro-
gen bonded polymer blends can be quite complicated”), but for our purposes here
all we need to know is that we are likely to observe significant mixing between PU
and PEO in the amorphous state. However, it is also important to recognize that
there are major differences in the phase behavior of segmented block SPUU copoly-
mers and the PUPPEO polymer blends discussed here. One cannot simply extrapo-
late the results obtained from polymer blends to predict the phase behavior, or type
and distribution of hydrogen bonds, that occur in segmented SPUU block copoly-
mers.
Fig. 14 shows the infrared spectra of PUPPEO blends containing 50 and 75 wt.-%
PEO, recorded at 80°C (above the T,,, of PEO) in the N-H and C=O stretching

5050
Pure PU
R u e PU
(Dim dered)
I ’
I ’

PurePU (Ordered) I ’
j i
3 b 35bo 34’00 33bo 3ux) 3100 1750 1700 1650 1600

W a v e n u m b e r in cm-’ W a v e n u m b e r in cm-’
Fig. 14. Scale expanded infrared spectra of PUPEO blend films containing 50 and
75 wt.-% PEO recorded at 80°C. Also included for comparison are the spectra of pure
PU in the ordered (“crystalline”) and disordered (amorphous) states. Left: the N-H
stretching region (3 100-3 600 cm-‘); right C=O stretching region ( 1 600- 1 750 cm-’)
Infrared temperature studies of a simple polyurea 135

regions. Also shown for clarity are the corresponding spectra of pure PU in the dis-
ordered (amorphous) and ordered (“crystalline”) states. To reiterate, in the N-H
stretching region the distinction between these two states is reflected in the overall
breadth of the N-H stretching band envelope, while in the carbonyl stretching
region three discrete C=O bands are observed attributed to (i) non-hydrogen bonded
or “free” (1 690 cm-’); (ii) disordered hydrogen bonded (1650 cm-’) and (iii)
ordered hydrogen bonded (1 630 cm-’) carbonyl groups.
In a miscible PUPEO blend, compositionally rich in PEO, the predominant
hydrogen bonded moiety will be a single PU urea group “capped” by an ether oxy-
gen of PEO, as illustrated below in Scheme 4:

Scheme L :

\ \
N-H
o=c: + o
\
(excess)

16dcm” /
3260 cm-’

This 1 : 1 uredether adduct would be expected to exhibit characteristic infrared


bands corresponding to “free” carbonyl groups at ~ 1 6 9 cm-’
0 and bifurcated hydro-
gen bonded N-H to ether oxygen groups at =3 200 to 3 300 cm-’. Examining the
spectra of the two PUPEO blends (Fig. 4) it is clear that the relative intensity of
both bands at 1690 and 3 260 cm-’ increase with increasing PEO concentration in
the blend and this is consistent with the above assignments (Scheme 4 ) .
As the concentration of PEO in the blend is reduced, however, equilibrium con-
siderations dictate that the contribution from disordered uredurea interactions
should increase. Indeed, the intensity of the contribution in the N-H stretching
envelope at 3 300 cm-’ is greater relative to that at 3 260 cm-’ for the 50 :50 com-
pared to the 25 :75 PUPEO blend. Similarly, the relative intensity of the “free” car-
bonyl band at 1690 cm-’ decreases, which is consistent with the expected increase
of hydrogen bonded uredurea carbonyl groups relative to non-hydrogen bonded car-
bonyl “end groups” (see Scheme 5 ) :

Scheme 5:

3300 cm-’ 3260 cm-’


136 M. M. Coleman, M. Sobkowiak, G . J. Pehlert, P. C . Painter, T. Iqbal

There is a shift in frequency of the band assigned to disordered uredurea hydro-


gen bonded carbonyl groups from ~ 1 6 5 cm-’ 0 in pure PU to ~ 1 6 5 cm-’
5 in the
P U P E O blends and this should not be brushed aside. If one compares the spectra of
the pure disordered PU to that of the 50: 50 P U P E O blend, it appears that the latter
still contains a contribution at 1650 cm-’. Accordingly, from the trends seen in the
spectra of the blends as a function of composition, we believe it is reasonable to ten-
tatively assign the 1665 cm-’ band to a hydrogen bonded uredurea carbonyl group
that is “capped” by an ether oxygen of PEO, as depicted in Scheme 5.

Acknowledgements: The authors thank Drs. H. L. Snyder and Z H. Kim of DuPont for
several stimulating conversations and gratefully acknowledge the financial support of the
National Science Foundation, Polymers Program.

’) (a) M. M. Coleman, D. J. Skrovanek, S. E. Howe, P. C. Painter, Macromolecules 18,


299 (1985);
(b) D. J. Skrovanek, S. E. Howe, P. C. Painter, M. M. Coleman, ibid. 18, 1676 (1985);
(c) D. J. Skrovanek, P. C. Painter, M. M. Coleman, ibid. 19,699 (1986);
(d) M. M. Coleman, D. J. Skrovanek, P. C. Painter, Makromol. Chem., Macromol.
Symp. 5,21 (1986)
2, M. M. Coleman, K. H. Lee, D. J. Skrovanek, P. C. Painter, Macromolecules 19, 2149
(1 986)
3, (a) M. M. Coleman, D. J. Skrovanek, J. Hu, P. C. Painter, Macromolecules 21, 59
(1988);
(b) M. M. Coleman, J. Hu, Y. Park, P. C. Painter, Polymer 29, 1659 (1988);
(c) J. Hu, P. C. Painter, M. M. Coleman, T. D: Krizan, J. Polym. Sci., Phys. Ed. 28,
149 (1990);
(d) D. E. Bhagwagar, P. C. Painter, M. M. Coleman, T. D. Krizan, J. Polym. Sci.,
Phys. Ed. 29, 1547 ( 1 99 1)
4, M. P. Stevens, Polymer Chemistry, 2nd ed., Oxford University Press, Oxford, England
1990
(a) H. Ishihara, I. Kimura, N. Yoshihara, J. Macromol. Sci., Phys. B22 ( 5 & 6), 713
(1 983-84);
(b) L. Born, H. Hespe, Coll. Polym. Sci. 263,335 (1985)
6, A. J. Ultee, “Fibers, Elastorneric”, in Encyclopedia of Polymer Science and
Engineering, vol. 6, 2nd ed., H. F. Mark, N. M. Bikales, C. G. Overberger and
G. Menges, Eds., J. Wiley & Sons, Inc., New York 1987, pp.733-755
7, (a) H. Ishihara, I. Kimura, K. Saito, H. Ono, J. Macromol. Sci., Phys. B10 (4), 591
(1974);
(b) C. S. Paik Sung, T. W. Smith, N. H. Sung, Macromolecules 13, 117 (1980);
(c) T. Yamamoto, M. Shibayama, S. Nomura, Polym. J. 21 (1 1), 895 (1989);
(d) Y. Xiu, Z. Zhang, D. Wang, S. Ying, J. Li, Polymer 33, 1335 (1992)
8, P. C. Painter, M. M. Coleman, “Fundamentals of Polymer Science”, Technomic Pub-
lishing, Inc., Lancaster, PA 1994
y, X. Yang, P. C. Painter, M. M. Coleman, Macromolecules 25,4996 (1992)
lo) (a) M. M. Coleman, D. F. Varnell, J. P. Runt, in: Contemporary Topics in Polymer
Science, vol. 4, W. J. Bailey, Ed., Plenum Press, New York 1981;
(b) D. F. Vamell, J. P. Runt, M. M. Coleman, Macromolecules 14, 1350 (1981)
‘I) (a) M. M. Coleman, J. F. Graft, P. C. Painter, “SpeciJiclnteractions and the Miscibil-
ity of Polymer Blends”, Technomic Publishing, Inc., Lancaster, PA, 1991;
(b) M. M. Coleman, P. C . Painter, Prog. Polym. Sci. 20, 1 (1995)

You might also like