You are on page 1of 9

Progress in Organic Coatings 54 (2005) 296–304

The phase mixing of moisture cured polyurethane-urea during cure


D.K. Chattopadhyay a , P.S.R. Prasad b , B. Sreedhar c , K.V.S.N. Raju a,∗
a Organic Coatings and Polymers Division, Indian Institute of Chemical Technology, Hyderabad 500007, India
b National Geophysical Research Institute, Hyderabad 500007, India
c Inorganic & Physical Chemistry Division, Indian Institute of Chemical Technology, Hyderabad 500007, India

Received 14 January 2005; accepted 8 July 2005

Abstract

Moisture cured polyurethane-ureas (MCPUs) is one of the industrially important polymer, which shows good thermal–mechanical and
weathering properties and widely used in the reactive hot melt adhesives and coatings. In this study, chemically crosslinked MCPUs were
prepared by reacting isophorone diisocyanate (IPDI) with polyethers like polytetramethyleneglycol (PTMG)-1000 and polyethyleneglycol
(PEG)-1000, with NCO/OH ratio 1.6:1. Trimethylol propane (TMP) was used as a crosslinking agent during the prepolymer synthesis. The
excess isocyanate of the prepolymers was cured with moisture at 25 ◦ C and humidity of 40%. Fourier transform infrared spectroscopy (FTIR)
and dynamic mechanical thermal analyzer (DMTA) measurements were used to monitor curing process of polyurethane-urea systems. Higher
correlation coefficient (R2 ) values were obtained for the second-order cure model compared to the first- and third-order for both the synthesized
prepolymers.

The change in short range ordering associated with hydrogen bonding as well as decrease in crystallinity of soft segment during the
phase mixing was observed from differential scanning calorimetry (DSC) measurements. The change in thermal stability was assessed by
thermogravimetric (TG) analysis. Characterizations of the curing process provide an essential base to obtain best polymer.
The phase mixing phenomenon was confirmed from the angle resolved X-ray photoelectron spectroscopy (AR-XPS).
© 2005 Elsevier B.V. All rights reserved.

Keywords: Polyurethane prepolymer; Moisture cure; Soft segment; Phase mixing

1. Introduction ings have been summarized by Gardner [1]. First of all,


they can be manufactured as a one-package system and their
Moisture-curing technology is one of the viable alterna- application is easier than the usual 2-K systems. Secondly,
tives in the use of low volatile organic component (VOC) since the salient reactant is water, the formulations have less
in the coating industry. This type of coating is one of VOC than 2-K polyurethane coatings. In comparison of 2-
the finest one available that can be produced without the K polyurethanes, the moisture-cured polyurethane/polyurea
application of heat or other external energy source. The coatings have good adhesion, abrasion resistance, thermal
advantages of moisture-curing polyurethane/polyurea coat- stability, hardness, chemical and solvent resistance, and high
humidity tolerance [2]. All of these properties are related
∗ Corresponding author.
and depend on the degree of cure. The excellent performance
E-mail addresses: kvsnraju@iict.res.in, drkvsnraju@yahoo.com results primarily from the formation of crosslinks during cure.
(K.V.S.N. Raju). The process of curing defines the properties of the finished

0300-9440/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.porgcoat.2005.07.004
D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304 297

coating material. Temperature, humidity level and the state other thermosetting materials, the cure of NCO-terminated
of cure are, therefore, key aspects for understanding and con- polyurethane prepolymer with moisture requires a care-
trolling the coating performance. Also the kinetics of cure ful control of processing conditions, in order to follow
allows one to develop a conversion profile at a particular an adequate modulus and strength development, which
temperature and relative humidity to assess the usefulness are strictly dependent on parameters such as relative
and relative importance of the different parameters leading humidity, catalyst type and concentration, and excess iso-
to the final product. cyanate content in the prepolymer. So, the determination
The MCPUs consist of isocyanate-capped low molec- of cure conditions for such type of coating system is a
ular weigh prepolymers prepared from polyester and or prime prerequisite for developing custom tailored properties
polyether glycols and diisocyanates. Water vapor from the [9–14].
atmosphere diffuses into the MCPU, and the nucleophilic In the present study, the focus is on the cure behav-
attack of water on NCO-terminated prepolymer results in ior and modulus development of two-moisture cure for-
an irreversible reaction, which produces carbamic acid. The mulations. Infrared spectroscopy was used for the quanti-
carbamic acid is unstable at room temperature and decom- tative evaluation of the extent of cure. Modulus measure-
poses into carbon dioxide and a primary amine (1). The ments were carried out by DMTA. Thermal stability and
primary amine is reactive with the NCO-terminated pre- calorimetric evaluation during cure was accomplished with
polymer and produces urea (2). This reaction leads to the TGA, and DSC instrument. The change in surface property
development of a three-dimensional network in the pres- during cure as well as with depth was evaluated through
ence of a tri-functional monomer in the reactive prepolymers AR-XPS.
[3,4]:
R NCO + H2 O → R NH2 + CO2 (↑) (1)

(2)
Crosslinking reaction are also possible between urea, ure-
thane and residual isocyanates end groups that results in the
formation of biuret (3) and allophanate (4) linkages.

(3)

(4)
However, Duff and Maciel [5] reported that, the predom-
inant postcure process is due to the reaction of isocyanate
group with atmospheric moisture to form an amine, which
further condenses with an additional isocyanate group from
the immediate vicinity to form a urea linkage. Cui et al. [6]
reported that the presence of side products enhances adhe- 2. Experimental
sion. Biuret and allophanate are more thermally stable than
the polyurethane/urea network as reported by Koscielecka 2.1. Materials
[7] and Kordomenos et al. [8].
In moisture cured polyurethane, the chemical reactions Poly(tetramethylene glycol) [PTMG, M.W. = 1000],
that take place within the material occur much faster than poly(ethylene glycol) [PEG, M.W. = 1000], trimethylol
the rate of water vapor diffusion. Thus, the material may propane (TMP), isophorone diisocyanate (IPDI) and dibutyl-
quickly form skins after application and exposure to the tin laurate (DBTL) from Aldrich (USA), 2-ethoxyethyl
air. This skin then acts as a barrier for the permeation of acetate, triethylamine and sulfur free toluene from S.D Fine
water into the remaining uncured material inside the bulk. chem. (Mumbai, India) were used. Solvents were stored over
Any water, which passes through the barrier layer quickly activated 3–4 Å molecular sieves.
reacts with the uncured material and thickness of the barrier
increases. 2.2. Method
The kinetic investigation is probably the most active
research because the rate and extent of crosslinking thor- The resin kettle was equipped with a dropping funnel,
oughly affect the mechanical properties. As in the case of stirrer, thermometer, reflux condenser and a nitrogen inlet.
298 D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304

Fifty grams PEG/PTMG and 9 g TMP mixture were reacted sentation output is produced by an integrated VISION control
with IPDI (with NCO:OH ratio of 1.6:1) at 75–85 ◦ C for 6 h and information system. All spectra presented are charge bal-
in absence of catalyst. These prepolymers were named as anced and energy referenced to C1s at 284.6 eV.
PEG/TMP/IPDI and PTMG/TMP/IPDI.

2.3. Sample preparation 3. Results and discussion

The catalyst (0.05% DBTL and 0.05% triethylamine) was 3.1. Cure behavior
mixed with the prepolymer and the films were cast on tin foil
by using a power driven automatic applicator. The supported The curing reaction involves a series of polyaddition reac-
films were amalgamated after 2–3 h. The underside of the tions, which depend on the curing conditions. During the
unsupported films was carefully cleaned with muslin cloth cure process of the formation of three-dimensional polyurea
and the free films were stored at 25 ◦ C and 40% RH. matrix, the interaction parameters between the polyether soft
and polyurea hard segments change with the reaction time and
2.4. Cure characterization lead to more and more microphase mixing from the initially
homogeneous state. Inside the bulk of the polymer, where
Infrared spectra of these films were recorded periodically the concentration of moisture is extremely low; in addition to
on NICOLET FTIR-740 Spectrophotometer with DTGS the formation of polyurea, the possibility exists for the for-
TEC detector with a resolution of 4 cm−1 and 128 scans. The mation of biuret and allophanate networks. Therefore, during
sample thickness was 0.15 mm. The crosslinking was fol- network maturation, a number of competitive reactions occur
lowed by monitoring the disappearance of isocyanate band. and the resultant morphology of the network is determined
The modulus development during network maturation was by two phenomena: the thermodynamic phase mixing and the
measured periodically using DMTA IV instrument (Rheo- kinetics of polymerization. Yanjun et al. [16] on the cure pro-
metric Scientific, USA) in tensile mode at a frequency of cess in a study of isocyanate-terminated urethane prepolymer
1 Hz and with a heating rate of 3 ◦ C/min by scanning films based on oligoester or oligoether diols and diisocyanates in
periodically from 35 to 100 ◦ C. Storage modulus (E ), loss saturated humidity level, shown that a second-order autocat-
modulus (E ) and tan δ as a function of temperature at a alytic model can be used to evaluate the curing process. The
constant frequency were obtained from these runs. Calori- authors have created a saturated steam environment inside
metric measurements were carried out using Mettler Toledo the DSC sample pan and by single run they evaluated the
DSC 821e thermal system, from temperature −60 to 200 ◦ C time needed for complete cure. In practice, the humidity level
at a heating rate of 10 ◦ C/min under nitrogen atmosphere is never 100% and therefore cure evolution is more com-
(flow rate, 30 ml/min). The instrument was calibrated with plicated. Therefore, in our study, the curing was followed
indium standards before measurements. The sample size was periodically by using the above techniques.
approximately 10 mg. The thermal stability of the polymer
film with the progress of cure was studied in N2 environment 3.2. FTIR spectroscopy
(flow rate, 30 ml/min) using Mettler Toledo TGA/SDTA 851e
thermal systems from 40 to 550 ◦ C and at a heating rate of The main aim of the present paper is to understand how
10 ◦ C/min. The sample size was 9–10 mg. The surface of the spectroscopic, thermal and thermo-mechanical changes
the sample was analyzed using a KRATOS AXIS 165 X-ray occur as the isocyanate groups are consumed during reaction
Photoelectron Spectrometer. The X-ray gun was operated at with moisture. Cure optimization is essential for such systems
15 kV voltage and 20 mA. Survey and high-resolution spectra as the rate of CO2 production during network growth may
were collected using 80 and 40 eV pass energy, respectively. adversely affect polymer property and decrease the cohesive
The pressure in the analyzer chamber was ∼1.33 × 10−6 Pa. forces between the hard segments. Excess humidity may cre-
A thin film of thickness 0.2 mm, without tin amalgamation ate premature curing, resulting in pinholes, foaming, or voids
was fixed on a carbon tape over a circular disc used inside the in the coating film. In this case, molecules may not orient in
XPS-chamber, and care was taken to record the spectra at the proper place or the system may lag the close packed struc-
same place every time. After each evaluation, the sample was ture as a result of less cure time. Again low humidity may
stored at the prescribed atmosphere until the second evalu- cause the polymer to take longer time for complete curing
ation was carried out. The sample was tilted in such a way with as result the probability that formation of side products
to change the angle θ between the normal to the sample and like biuret and allophanate increases inside the bulk of the
the analyzer. At θ = 0◦ , the sample was perpendicular to the polymer [17,18]. In the present investigation, RH was main-
detector, leading to the maximum sampling depth. The effec- tained at 40% and the temperature was 25 ◦ C. The isocyanate
tive sampling depth, z, was derived by z = 3λ cos θ, where λ absorption band occurs at approximately 2300–2270 cm−1
is the effective mean free path for electrons to escape the sur- in the mid-infrared spectrum, and the decay in the intensity
face and was set to the value of 2.5 nm. Therefore, at θ = 0◦ , of this absorbance was used to monitor the conversion of
z = 7.5 nm and at θ = 80◦ , z = 1.3 nm [15]. Peak fitting and pre- isocyanate-functional groups during the curing reaction. In
D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304 299

Fig. 1. (a) The representative FTIR spectra of PTMG/TMP/IPDI, shows the decay of NCO group (film thickness = 0.15 mm). (b) The integrated NCO peak
area vs. time for PTMG/TMP/IPDI and PEG/TMP/IPDI.

Fig. 1a, the change in spectral intensity (disappearance of A basic kinetic equation is
NCO group) of PTMG/TMP/IPDI MCPU with cure time is
dp
shown. From 70 h (trace 1) to 1054 h (trace 7), it was clearly = k(T )f (p), (7)
observed that the polyurethane was close to being fully cured. dt
This is evidence by the drastic reduction of the NCO peak at and when presented in isothermal conditions
2273 cm−1 [3]. Fig. 1b represents the change in NCO group  
dp −E
spectra intensity with cure time for the two systems investi- = A exp f (p), (8)
dt RT
gated, and showed an asymptotic trend of decay of the NCO
concentration [19]. From the slope of the Fig. 1b, we suggest where A is the pre-exponential factor and E is the activation
that the cure rate of PEG/TMP/IPDI was slower in compari- energy of the curing process. Two empirical schemes, nth-
son to PTMG/TMP/IPDI system. order and autoaccelerated models, are widely used for mod-
In order to obtain the degree of curing it is assumed that eling the cure kinetics for thermosetting materials [22,23].
there are no side reactions. Hence, the isocyanate conversion The nth-order kinetics can be expressed as
can be used as the degree of curing as follows:
dp
= k0 (1 − p)n . (9)
At − A ∞ dt
Isocyanate conversion (p) = 1 − (5)
A0 − A ∞ and the autoaccelerated model as
Here, A0 is the normalized area of the absorption at the initial dp
= (k1 + k2 pm )(1 − p)n , (10)
time, At the normalized area of the absorption at a certain time dt
during the curing process, and A∞ is the final normalized Here, k0 , k1 , and k2 are constants related to the rate constants,
area of the absorption at infinite time. For a completely cured which depend on temperature. Parameters m and n are related
system A∞ will be zero, because no NCO functionality will to the reaction order. Here, we used the nth-order approach
be available for IR absorption in such case. for modeling the curing kinetics of MCPUs:

3.3. Crosslinking kinetics If n = 1, ln(1 − p) = −k0 t + C (11)


1
If n = 2, = k0 t + C (12)
Chemical reactions that take place during the curing reac- 1−p
tion of a thermoset determine the polymer morphology, which
1
is a key factor in determining the properties of the cured ther- If n = 3, = k0 t + C (13)
moset. Consequently, the understanding of the mechanism (1 − p)2
and kinetics of the curing is very important in the evalua- The fitting results are summarized in Table 1. To com-
tion of the structure–morphology–property relationships of a pare the validity of first-order, second-order and third-order
material. The degree of cure p can be expressed as: models, linear regression treatments were performed on the
 t investigated systems. The correlation coefficients (R2 ) in the
p(t) = dp (6) linear regression analysis were used to evaluate the validity
0 of the kinetic models. For both systems, higher R2 values
Based on empirical rate laws, a variety of kinetic mod- were obtained for the second-order modeling. This demon-
els [20,21] can be adopted to describe the curing process of strates that the overall curing process favored a second-order
thermosetting systems. reaction [24].
300 D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304

Table 1
Rate constants and fitting results for the curing process of MCPUs derived
from FTIR measurements (p, range of curing degree fitted to the reaction
order; k, rate constant (h); R2 is the correlation coefficient of the fit)
System FTIR First-order Second-order Third-order
PTMG/TMP/IPDI p 0–0.87 0–0.97 0–0.97
k 0.004 0.006 0.0268
R2 0.92 0.97 0.92
PEG/TMP/IPDI p 0–0.96 0–0.97 0–0.97
k 0.0024 0.0015 0.0055
R2 0.93 0.96 0.89

3.4. Dynamic mechanical analysis

The elastic storage modulus, E versus temperature and


the loss factor versus temperature spectra of representative Fig. 3. Modulus (E ) vs. time plot of PTMG/TMP/IPDI and PEG/TMP/IPDI,
shows the increase of modulus due to polymer network growth.
film PTMG/TMP/IPDI with different cure times are shown in
Fig. 2a and b, respectively. E and E characterize the elastic
more than 99% after 1000 h. Therefore, from the spectrum
and viscous component of a material under deformation, E
of E versus temperature E recorded at maximum cure time
is a measure of mechanical energy stored under load.
was taken as E at t = ∞:
The E value at 45 ◦ C for PTMG/TMP/IPDI system after
  
46 h of cure was 8 × 106 Pa, which changed to 3.7 × 108 Pa Et Et
after 1390 h. Similarly, the E value also increased consid- p =  or  (14)
E∞ E∞
erably for PEG/TMP/IPDI system. The change in E value
with cure time is much more noticeable in the glassy plateau
region of the DMTA profile. The change of modulus at 45 ◦ C 3.5. Isocyanate conversion
with time depicts the change in crosslink density and molec-
ular weight build up due to crosslinking. Initially, the casted Eqn. (11)–(13) were used to find the rate constant. The
films after complete solvent evaporation were in soft con- representative plots are shown in Fig. 4, and the rate constant
dition with low storage modulus (E ) at 45 ◦ C. With time, data are reported in Table 2. For both the systems, higher
films developed strength as a result of NCO and moisture R2 values were obtained for the second-order modeling, and
reaction. Similarly, E also increases during the network mat- therefore, the overall curing process favored a second-order
uration as shown in Fig. 2b. It was observed that the modulus reaction.
buildup of PTMG/TMP/IPDI was more than PEG/TMP/IPDI
after certain period of time (Fig. 3). 3.6. DSC analysis
From these observations, a correlation of isocyanate con-
version with cure time t was made. Isocyanate conversion at Fig. 5 shows the DSC thermogram of the investigated sam-

cure time t was the modulus (Et or Et ) at that time divided ples. The soft segment glass transition temperature, TgS , of
by the modulus at time t = ∞. It was observed from FTIR PEG/TMP/IPDI system observed initially at −33 ◦ C, disap-
spectral measurements that, the isocyanate conversion was peared within a few hours of cure of the polymer (Fig. 5a).

Fig. 2. (a) E vs. temperature plot of PTMG/TMP/IPDI film showed the change of modulus with cure time, film thickness was 0.15 mm. (b) E vs. temperature
plot of PTMG/TMP/IPDI film.
D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304 301

3 h after casting the film was 2.8 J g−1 , which changes to


43.1 J g−1 after 2 months of cure (Table 3). At the same
time the Tm value changed from 63.8 to 78.2 ◦ C. Similar
phenomenon was also observed for the PTMG/TMP/IPDI
system. These results showed the enhancement in the short
range ordering with the network maturation, and demonstrate
that, crystalline structure will diminish with moisture cure
[25].

3.7. TGA analysis

The weight loss versus temperature plot of


PEG/TMP/IPDI and PTMG/TMP/IPDI systems with
different degree of cure is shown in Fig. 6a and b, respec-
tively. Both the coatings showed a two step decomposition
pattern, and are thermally stable upto 250 ◦ C. The character-
istic thermal data obtained from the TG measurements with
different cure time are reported in Table 3. The onset decom-
position temperature (Ti1 ) for the first decomposition step of
PEG/TMP/IPDI after 10 days was 268 ◦ C, which changes
to 278 ◦ C after 2 months. Similarly, for PTMG/TMP/TDI,
the Ti1 value changes from 265 to 288 ◦ C from 3 h to 2
months of cure. Therefore, considerable enhancement in
the thermal stability with the progress of cure was achieved
for both systems. The Ti1 value after 2 months of cure
for PTMG/TMP/IPDI system was 10 ◦ C more than for the
PEG/TMP/IPDI system. Similarly, substantial improvement
in the decomposition onset temperature for the second stage
Fig. 4. Relationship between p and curing time of polyurethane/ureas deter- (Ti2 ) and endset decomposition (Tf ) temperatures were
mined from E values (DMTA). also observed. The corresponding DTG curve (not shown)
shows a high temperature shift of maximum decomposition
temperatures (Tmax ) for both the degradation stages. The
The endotherm at the soft segment Tg was related to its crys-
char yield value at 450 ◦ C, suggests that at high temperature
tallinity. The decrease and disappearance of this endotherm
PEG/TMP/IPDI was more stable than PTMG/TMP/IPDI.
suggests the reduction in soft segment crystallinity, which is
The observable difference between the investigated systems
due to phase mixing. A broad endotherm from 40 to 120 ◦ C in
is that, the second degradation step is more separated from
the DSC thermogram appeared, and was associated with the
the first decomposition step for PTMG/TMP/IPDI, whereas
short range order. This observation suggests that with cure
not much separation between these two steps is observed for
time the enthalpy associated with the breakdown of hydro-
PEG/TMP/IPDI.
gen bonding between hard segment polyurethane/urea groups
and the ether-oxygens of the soft segment increases and
broadening of the transition takes place. The minima in the 3.8. XPS analysis
endotherm, Tmin , shifted towards the high temperature region
and transition breadth increases with the progress of cure. For In order to confirm the phase mixing behaviour, we
example, the enthalpy value (H) of the PEG/TMP/IPDI have recorded high resolution as well as narrow scans for

Table 2
Rate constants and fitting results for the curing process of MCPUs derived from DMTA measurements (p, range of curing degree fitted to the reaction order; k,
rate constant (h); R2 is the correlation coefficient of the fit)
System DMTA First-order (E /E ) Second-order (E /E ) Third-order (E /E )
PTMG/TMP/IPDI p 0.02–0.81/0.03–0.68 0.02–0.81/0.03–0.68 0.02–0.81/0.03–0.68
k 0.0018/0.0012 0.0045/0.0022 0.027/0.0088
R2 0.97/0.96 0.98/0.99 0.91/0.97
PEG/TMP/IPDI p 0.02–0.44/0.01–0.21 0.02–0.44/0.01–0.21 0.02–0.44/0.01–0.21
k 0.0015/0.0006 0.002/0.0007 0.0056/0.0016
R2 0.96/0.98 0.97/0.993 0.96/0.99
302 D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304

Fig. 5. (a) The change in Tm and Cp with the cure time for PEG/TMP/IPDI and (b) change in Tm and H with the cure time for PTMG/TMP/IPDI. Down
sided arrows shows the temperature corresponds to minima (Tm ) on the endothermic peak, while up sided arrows showed glass transition temperature.

Table 3
The TGA and DSC data of PEG/TMP/IPDI and PTMG/TMP/IPDI with different cure time
Sample name Cure time Ti1 (◦ C) T1max (◦ C) Ti2 (◦ C) T2max (◦ C) Tf (◦ C) wt.% at 450 ◦ C H (Jg−1 ) Tm
PEG/TMP/IPDI 3h 261 305 336 362 389 1.0 2.8 63.8
10 d 268 312 340 368 393 2.2 15.7 64.0
15 d 273 314 343 372 394 3.0 24.8 73.7
2m 278 318 345 375 397 7.8 43.1 78.2
PTMG/TMP/IPDI 3h 265 310 340 378 415 1.2 5.4 65.2
10 d 278 313 355 383 417 1.8 25.7 68.0
15 d 283 317 360 387 420 2.8 34.9 74.0
2m 288 322 366 388 425 3.2 46.3 81.4
h, hours; d, days; m, months.

Table 4
AR-XPS narrow scan atom percent data of PTMG/TMP/IPDI at takeoff angle 0 and 80◦ with different cure time
Cure time C (0◦ /80◦ ) N (0◦ /80◦ ) O (0◦ /80◦ ) C O/C C ratio (0◦ /80◦ )
2h 78.49/79.45 4.22/3.10 17.3/17.45 0.214/0.403
8h 82.53/72.24 3.12/8.00 14.35/19.76 0.365/0.307
10 d 79.10/68.19 2.90/17.21 18.00/14.60 0.519/0.287

PTMG/TMP/IPDI with different cure time. The peaks at 531, As nitrogen is associated with the urethane and urea link-
400, and 285 eV due to oxygen (O1s), nitrogen (N1s), and car- ages, the XPS data (Table 4) indicate a significant decrease in
bon (C1s), respectively, were observed in the survey spectra nitrogen concentration from the bulk to surface, which signi-
at takeoff angle 0◦ and 80◦ . The observed XPS analysis data fies that the urethane and urea content at the upper surface is
in Table 4 corresponds to the sample depth of 7.5 and 1.3 nm. much less than in the bulk (after 2 h) [26]. The chemical struc-

Fig. 6. (a) The thermogravimetric curves of PEG/TMP/IPDI and (b) PTMG/TMP/IPDI with cure time.
D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304 303

Fig. 7. The C1S AR-XPS figures of PTMG/TMP/IPDI at 0◦ and 80◦ with different cure time, showed the phase mixing character during network maturation.

ture of PTMG/TMP/IPDI system is complicated and contains ure not shown) are consistent overall with the C1s profiles,
a number of carbon atoms in different environments with and are comprised mainly of C O (530.5–532.5 eV) and C O
their different binding energies. Again C1s bonding ener- (532–534 eV). This experimental investigation confirmed the
gies of these functional groups were very close to each other segregation behavior of soft segments towards the surface,
and difficult to resolve in the curve fitting analysis. There- and was associated with the surface energy, the surface being
fore, these groups were combined and classified into three enriched with lower energy components.
component groups (Fig. 7). One corresponding to aliphatic The increase in C O/C C ratio at 0◦ and its decrease
or cycloaliphatic carbon atoms bound to hydrogen or to car- at 80◦ takeoff angles with the progress of cure suggest that
bon atom appearing at 284.4–284.8 eV (C C and C H). The phase mixing takes place during the network maturation. The
peak associated with the etheral carbon (C O) contribution enhancement of microphase mixing provides a lower driv-
from the PTMG soft segment (285.4–286.2 eV) and the peak ing force for soft segments to localize on the surface. These
corresponding to carbonyl carbon (C O), from urethane and phenomena will also be intensified by the lowering in the
urea groups and appearing at 288.0–289.0 eV. After 2 h of mobility of the soft segments with the cure time. Therefore,
curing of the PTMG/TMP/IPDI system, an increase of C O microphase mixing affects the surface properties of MCPUs
peak percentage with increasing takeoff angle (i.e. towards [29].
surface) was observed, which suggests that the ether groups
preferentially resided towards the top surface of the polymer.
Since the low-resolution XPS data (Table 4) indicate a deple- 4. Conclusion
tion of N-containing groups and therefore an enrichment of
PTMG soft segments on the surface, it is suggested that the Infrared spectroscopy and DMTA analyses were used to
increase of C O/C C ratio on the surface would reflect in monitor the concentration changes in isocyanate groups dur-
part the reduction of the urethane linkages at the beginning ing polyurethane curing and to follow modulus development.
of cure [27–28]. The high-resolution O1s XPS spectra (fig- The reaction kinetics and the rate constant of the cure process
304 D.K. Chattopadhyay et al. / Progress in Organic Coatings 54 (2005) 296–304

were evaluated through a model that shows a second-order [7] A. Koscielecka, Acta Polym. 42 (1991) 221.
reaction. The change in sample short-range ordering enthalpy [8] P.I. Kordomenos, J.E. Kresta, K.C. Frisch, Macromolecules 20
and the stability with cure time were assessed from DSC (1987) 2077.
[9] R. Lomolder, F. Plogmann, P. Speier, J. Coat. Technol. 69 (868)
and TGA measurements. The observation suggests that the (1997) 51.
rate of cure was different for the two systems investigated [10] S.D. Seneker, T.A. Potter, J. Coat. Technol. 63 (793) (1991) 19.
and results in different reaction rate constant. The degree [11] B.W. Ludwig, M.W. Urban, J. Coat. Technol. 68 (857) (1996)
of phase mixing after some period of time was found to 93.
depend on the soft segment structure and created different [12] G.N. Robinson, J.F. Alderman, T.L. Johnson, J. Coat. Technol. 65
(820) (1993) 51.
stability. The thermal stability data suggests that the inves- [13] A.M. Heintz, D.J. Duffy, S.L. Hsu, Macromolecules 36 (2003) 2695.
tigated samples were stable upto 250 ◦ C without any major [14] R. Narayan, D.K. Chattopadhyay, B. Sreedhar, K.V.S.N. Raju, J.
decomposition and that PTMG/TMP/IPDI was more stable Mater. Sci. 37 (2002) 4911.
than PEG/TMP/IPDI. Higher R2 values were obtained for [15] Y. Deslandes, G. Pleizer, D. Alexander, P. Santerre, Polymer 39
(1998) 2361.
the second-order modeling than for the first or third-order
[16] C. Yanjun, H. Ling, W. Xinling, T. Xiaozhen, J. Appl. Polym. Sci.
modeling for both the synthesized prepolymers. The phase 89 (2003) 2708.
mixing phenomenon was confirmed from the AR-XPS mea- [17] B.W. Ludwig, M.W. Urban, J. Coat. Technol. 66 (839) (1994)
surements. 59.
[18] A.M. Kaminski, M.W. Urban, J. Coat. Technol. 69 (837) (1997) 113.
[19] G. Malucelli, A. Priola, F. Ferrero, A. Quaglia, M. Frigione, C.
Carfagna, Int. J. Adhesion Adhesives 25 (2005) 87.
Acknowledgement [20] S. Yi, H.H. Hilton, J. Comput. Mater. 32 (1998) 600.
[21] N. Lisardo, F. Fraga, L. Fraga, A. Castro, J. Appl. Polym. Sci. 63
D.K. Chattopadhyay acknowledges the receipt of an Uni- (1997) 635.
versity Grants Commission (UGC, India) Award for Research [22] Y. Liu, X.D. Sun, X.Q. Xie, D.A. Scola, J. Appl. Polym. Sci. 36
Fellowship in Engineering & Technology. (1998) 2653.
[23] J.W. Chen, L.W. Chen, J. Polym. Sci: Part A: Polym. Chem. 36
(1998) 3073.
[24] S. Li, R. Vatanparast, H. Lemmetyinen, Polymer 41 (2000) 5571.
References [25] Y. Cui, D. Chen, X. Wang, X. Tang, Int. J. Adhesion Adhesives 22
(2002) 317.
[1] G.J. Gardner, J. Prot. Coat. Linings 13 (2) (1996) 81. [26] F.Z. Sidouni, N. Nurdin, P. Chabrecek, D. Lohmann, J. Vogt, N.
[2] H. Ni, A.D. Skaja, M.D. Soucek, Prog. Org. Coat. 40 (2000) 175. Xanthopoulos, H.J. Mathieu, P. Francois, P. Vaudaux, P. Descounts,
[3] J. Comyn, F. Brady, R.A. Dust, M. Graham, A. Haward, Int. J. Surf. Sci. 491 (2001) 355.
Adhesion Adhesives 18 (1998) 51. [27] G.B. Wang, R.S. Labow, J.P. Santerre, Macromolecules 33 (2000)
[4] J. Comyn, Int. J. Adhesion Adhesives 18 (1998) 247. 7321.
[5] D.W. Duff, G.E. Macial, Macromolecules 24 (1991) 387. [28] L. Sabbatini, P.G. Zambonin, J. Electron Spectr. Relat. Phenomena
[6] Y. Cui, D. Chen, X. Wang, X. Tang, Int. J. Adhesion Adhesives 22 81 (1996) 285.
(2002) 317. [29] K. Nakamae, T. Nishino, S. Asaoka, Sudaryanto, Int. J. Adhesion
Adhesives 16 (1996) 233.

You might also like