Anda di halaman 1dari 13

Attenuated Total Reflectance/Fourier Transform Infrared

Studies on the Phase-Separation Process of Aqueous


Solutions of Poly(N-isopropylacrylamide)

ORY RAMON,1 ELLINA KESSELMAN,1 RONEN BERKOVICI,2 YACHIN COHEN,2 YARON PAZ2
1
Department of Food Engineering and Biotechnology, Technion, Haifa 32000 Israel

2
Department of Chemical Engineering, Technion, Haifa 32000 Israel

Received 6 February 2001; revised 9 April 2001; accepted 27 April 2001


Published online 00 Month 2001

ABSTRACT: Temperature-induced phase separation of poly(N-isopropylacrylamide) in


aqueous solutions was studied by attenuated total reflectance (ATR)/Fourier transform
infrared spectroscopy. The main objectives of the study were to understand, on a
molecular level, the role of hydrogen bonding and hydrophobic effects below and above
the phase-separation temperature and to derive the scenario leading to this process.
Understanding the behavior of this particular system could be quite relevant to many
biological phenomena, such as protein denaturation. The temperature-induced phase
transition was easily detected by the ATR method. A sharp increase in the peaks of both
hydrophobic and hydrophilic groups of the polymer and a decrease in the water-related
signals could be explained in terms of the formation of a polymer-enriched film near the
ATR crystal. Deconvolution of the amide I and amide II peaks and the OOH stretch
envelope of water revealed that the phase-separation scenario could be divided, below
the phase-separation temperature, into two steps. The first step consisted of the
breaking of intermolecular hydrogen bonds between the amide groups of the polymer
and the solvent and the formation of free amide groups, and the second step consisted
of an increase in intramolecular hydrogen bonding, which induced a coil– globule
transition. No changes in the hydrophobic signals below the separation temperature
could be observed, suggesting that hydrophobic interactions played a dominant role
during the aggregation of the collapsed chains but not before. © 2001 John Wiley & Sons,
Inc. J Polym Sci Part B: Polym Phys 39: 1665–1677, 2001
Keywords: Fourier transform infrared (FTIR); hydrogen bonding; poly(N-isopro-
pylacrylamide); phase separation

INTRODUCTION mational changes such as coil– globule, coil– helix,


and association to supramolecular assembly.1 In
The solution properties of many nonionic water- a coil– globule transition (CGT), an expanded flex-
soluble polymers are temperature-dependent and ible polymer chain turns into a collapsed globule2
exhibit phase separation of the lower critical so- that precipitates on aggregation. The CGT phe-
lution temperature (LCST) type. During the pro- nomenon has been studied extensively both theo-
cess, the macromolecules exhibit dramatic confor- retically and experimentally.3 Several studies
concentrated on polystyrene in various organic
solvents,4 determining the molecular parameters
Correspondence to: Y. Paz (E-mail: paz@tx.technion.ac.il)
related to the initial and final states, without
Journal of Polymer Science: Part B: Polymer Physics, Vol. 39, 1665–1677 (2001)
© 2001 John Wiley & Sons, Inc. clarifying in detail the crucial role of intermolec-
1665
1666 RAMON ET AL.

ular and intramolecular interactions in polymer– pic ratio have led to a novel description of the
solvent systems. structure and dynamic of water at a molecular
An investigation of CGT in water-soluble syn- level.20 The same technique was found to be ben-
thetic polymers can be very useful in elucidating eficial for extracting information on polymers,
the mechanisms governing denaturation (folding) proteins, and their interactions with water.21 In
of large proteins.5 A common model polymer for two pioneering works, ATR/Fourier transform in-
such studies is poly(N-isopropylacrylamide) frared (FTIR) was used in studies related to the
(PNIPA), a polyvinyl polymer, containing hydro- temperature-dependent phase separation of
philic amide groups and hydrophobic isopropyl PNIPA gels22 or PNIPA–water solutions.23 In the
groups.6 Because of its side chains, the polymer is latter study, relevant information at a molecular
soluble in polar solvents (e.g., water) and low- level was derived on the role of intermolecular
polarity solvents such as tetrahydrofuran.5 The and intramolecular hydrogen bonding and the en-
crosslinked PNIPA gel in water undergoes a re- hancement of the hydrophobic interactions in this
versible volume-phase transition due to changes system.
in temperature and solvent composition.7 These In this study, the ATR technique was used to
thermoreversible phase-separated systems were investigate the phase separation of well-defined
studied for their potential use as intelligent ma- aqueous PNIPA solutions. The main objective
terials in controlled drug delivery, enzyme immo- was to clarify the mechanisms leading to the
bilization, and separation processes.8 It has been phase-separation phenomenon by interpretation,
proposed9 that phase separation in PNIPA solu- on a molecular level, of the role of hydrogen bond-
tions is a two-step mechanism, initiated by the ing and hydrophobic effects below and above the
collapse of single chains and advanced by aggre- phase-separation point and by elucidation of the
gation. According to this explanation, PNIPA role of water in the phase-transition process.
chains in aqueous solutions form expanded coils
at room temperature that undergo a CGT when
the temperature is increased to 32–35 °C (ca. the EXPERIMENTAL
␪ temperature). The relative roles of hydrogen
bonding10 and hydrophobic interactions11 in pro- Preparation of PNIPA
moting the phase transition are a matter of some
debate. PNIPA samples were prepared with a redox poly-
Although the gross scenario of PNIPA solution merization in an aqueous solution at 4 °C with a
phase separation is well understood, the detailed protocol similar to that used by Otake et al.24 The
scenario of the process is still under debate. More reagents were N-isopropylacrylamide (NIPA)
specifically, the role of water–PNIPA hydrogen monomer (Aldrich Chemical Co., United States),
bonding versus the role of intramolecular hydro- ammonium persulfate (APS; Merck, Darmstadt,
gen bonding is not yet fully understood. Many Germany), and N,N,N⬘,N⬘-tetramethylethyl-
methods, including cloud-point determination,12 enediamine (TEMED; Aldrich Chemical). The
light scattering,13 thermal analysis, differential NIPA monomer was recrystallized from a 1:1 tol-
scanning calorimetry (DSC) and microcalorim- uene/petrol-ether mixture as described else-
etry,14 rheology,15 fluorescence,16 and NMR,17 where.25 The polymerization was conducted ac-
have been used to study the phase-separation cording to the following procedure: 0.689 mL of
phenomena. However, these methods lack the NIPA monomer was dissolved in doubly distilled
ability to provide detailed information on a mo- water, and 2 mL of 4% APS was dissolved sepa-
lecular level. Pioneering studies of aqueous poly- rately. Then, the two solutions were mixed and
mer solutions by direct absorption IR spectrosco- cooled to 4 °C, and nitrogen was bubbled over the
py18,19 suggest that the water structure around a solution to remove oxygen. At that stage, 0.48 mL
solute is different than that in bulk water. How- of TEMED was added, and polymerization contin-
ever, such measurements are hampered at the ued for 3 h under a nitrogen atmosphere. The
fundamental transitions by the exceptionally polymer solution was then dialyzed for 6 days at 4
high absorbency of water. Implementation of the °C [the molecular cutoff was a weight-average
attenuated total reflectance (ATR) setup has molecular weight (Mw) of 14,000] to remove non-
given a way to overcome this obstacle. Quantita- reacted monomers, initiator, and catalyst, and
tive analyses of changes in these spectra with the then the unfractionated polymer was freeze-dried
temperature or Hydrogen/Deuterium (H/D) isoto- (Martin Christ, Alpha 1, Germany). The molecu-
ATR/FTIR STUDIES OF PNIPA 1667

Figure 1. DSC peak of an aqueous solution of PNIPA (2 wt %).

lar weight (7 ⫾ 1 MDa) was evaluated from the Bruker IFS55 machine equipped with a temper-
determination of the intrinsic viscosity of the ature-controlled circle cell accessory [circle cell
PNIPA aqueous solutions with an Ubbelohde cap- attenuated total reflection (CCATR)], as shown in
illary viscometer (at 20 °C) and the expression [␩] Figure 2. The accessory was made of a ZnSe cy-
⫽ 0.112 Mw0.51 (cm3/g), developed by Kubota et lindrical crystal (75 mm, 6-mm diameter, 45° edg-
al.5,7 es), fixed by a leak-proof O-ring at the center of a
temperature-controlled stainless steel vessel (60
mm long, 15-mm inner diameter) containing the
Determination of the Phase-Transition
PNIPA solution. To increase the signal/noise ra-
Temperature
tio, the IR beam was first decollimated and then
The phase-transition temperature was deter- redirected onto the crystal edges in a circular
mined by two thermal analysis techniques. The manner that enabled the probing of the entire
first, DSC, was done with a PerkinElmer DSC-7 circumference of the embedded crystal, with a
apparatus at a scanning rate of 5 °C/min. Figure reflecting angle of 45°. The IR beam that im-
1 presents the thermogram of a 2% (w/w) aqueous pinged on the edges of the crystal, outside of the
solution of PNIPA. The figure shows an endother- vessel, propagated through the crystal and exited
mic peak (0.81 J/g) at 36.2 °C, with an onset at through the other end, where it was redirected by
35.1 °C. The second technique, microcalorimetry, another set of mirrors toward the IR detector. The
was performed with a VP-DSC microcalorimeter short distance from which the signal was collected
(Micro-Cal Inc., Northampton, MS). Solutions of by this method (a few micrometers) minimized
0.5 wt % PNIPA in water were degassed and problems related to the high absorption coeffi-
transferred to a sample cell (0.5 mL) with a cali- cient of water. The temperature of the sample
brated syringe. The thermograms were obtained solution was monitored by a thermocouple in-
at heating rates of 1.5 °C/min. Polymer-free solu- serted into the vessel. The ATR/FTIR spectra
tions (water) were used in the reference cell. were taken in the range of 20 – 48 °C with a res-
olution of 1 °C. The circle cell was equipped with
a transparent poly(methyl methacrylate) cover,
ATR Spectroscopy Studies
allowing visual observation of the temperature-
ATR/FTIR measurements of PNIPA aqueous so- induced transition from a clear solution to a tur-
lutions (1–3.5 wt %) were performed with a bid solution.
1668 RAMON ET AL.

Figure 2. CCATR/FTIR setup.

All measurements were taken with the empty band peaking at 2125 cm⫺1, a peak that is known
circle cell as a background. Because the IR spec- to be a good indication for water quantity.27 An-
trum of water is known to change as a function of other reason for using this peak as a subtraction
temperature, mainly because of the breaking of
hydrogen bonds,26 care was taken to record a set
of temperature-dependent spectra of water
(20 – 48 °C), later to be subtracted from the tem-
perature-dependent spectra of the aqueous poly-
mer solutions.

RESULTS

Phase-Separation Effect as Probed by the PNIPA IR


Peaks
Figure 3 presents the ATR raw spectra of a 2.5%
aqueous solution of PNIPA at 22, 34, 35, 36, 37,
and 45 °C. As mentioned in the Experimental
section, the background here was an empty cell,
so the spectra are dominated by the water signals
and, more specifically, the OOH stretch envelope
of the solvent at 3000 –3700 cm⫺1, the HOH bend-
ing mode at 1635 cm⫺1, and the wide combination
band peaking at 2125 cm⫺1. At temperatures
greater than 35 °C (the phase-separation temper-
ature as measured by DSC; Fig. 1), a series of
peaks seem to emerge in the spectra. These peaks
could easily be identified as those of the polymer.
To obtain better insight into the temperature-
induced changes in the spectra, we subtracted the
spectra of water at the corresponding tempera-
tures from that of the raw data. To account for the
effects of dilution, changes in penetration depth, Figure 3. Spectra of a 2.5% PNIPA solution at six
and interaction between the solvent and the poly- different temperatures. The polymer signals can hardly
mer, we fitted a multiplication constant to the be resolved because of the dominance of the water-
water spectra. Fitting was based on obtaining a related peaks. Note the difference between the 35 °C
perfect elimination of the water wide combination spectrum and the 36 °C spectrum.
ATR/FTIR STUDIES OF PNIPA 1669

2882 to 2878 cm⫺1, the CH2(a) peak shifted from


2939 cm⫺1 to 2936, and the CH3(a) peak shifted
from 2982 to 2975 cm⫺1. Line-shape fitting of
these peaks revealed a gradual decrease in their
width below the phase-transition point, followed
by an increase in their width at the phase-sepa-
ration temperature. Above the phase-separation
point and after oversubtraction of water, another
three weak peaks could be resolved. These were
the amide II overtone at 3070 cm⫺1, the hydro-
gen-bonded NOH stretch at 3310 cm⫺1 and an-
other peak at 3550 cm⫺1 (free NOH stretch?)
with intensities of 5, 10, and 7% of that of the
amide II, respectively.
The qualitative description of Figure 4 is given
in a more quantitative manner in Figure 5. Here,
the absorbance, integrated over the envelope of
each peak, is presented as a function of tempera-
ture. According to the figure, the temperature
dependency of the integrated absorbance of all
peaks could be divided into four regions:

1. Below 31 °C, a region of constant intensity.

Figure 4. Subtraction spectra of a 2.5% PNIPA solu-


tion at six different temperatures obtained by the sub-
traction of the water contribution from the crude spec-
tra presented in Figure 3.

guide was the fact that this spectral range is clear


of polymer-related peaks, unlike, for example, the
HOH bending peak that partially overlaps the
amide I peak of PNIPA. For a 2.5% solution, the
multiplication factor below the phase-transition
temperature ranged between 0.98 and 0.99,
whereas above the phase-separation point, the
value of this factor was no more than 0.94 – 0.95.
The subtraction spectra are presented in Figure
4. The phase transition is clearly indicated by the
intensity enhancement in the amide I (mostly
CAO stretch) band and amide II (mostly NOH
in-plane deformation) band at 1628 and 1556
cm⫺1, respectively,28 and in the coupled band of
COH and CH3 at 1460 cm⫺1. In parallel, an in-
crease in the signal of the isopropyl group22 at
1355–1400 cm⫺1 was observed, together with that
Figure 5. Changes in the integrated intensity of the
of the CH2 antisymmetric mode and the CH3 sym- PNIPA peaks as a function of temperature: (■) amide I
metric and antisymmetric stretching modes. (1628 cm⫺1), (Œ) amide II (1556 cm⫺1), (F) CH3 and
Here, a sharp change in the location of the peaks CH2 stretch envelope (2850 –3000 cm⫺1), (䊐) isopropyl
toward lower frequencies was observed with the I (1418 –1484 cm⫺1), and (E) isopropyl II (1358 –1418
phase transition: the CH3(s) peak shifted from cm⫺1).
1670 RAMON ET AL.

Figure 6. Deconvoluted spectra of the amide I and amide II peaks at (A) 25 and (B)
40 °C, showing the intermolecular, intramolecular, and free components of each band.

2. At 31–35 °C, a pretransition region in bonded subband at 1535 cm⫺1. These wave num-
which the intensity of some of the peaks is bers are in line with the notion that hydrogen
reduced slightly, whereas the intensity of bonding tends to make steeper the potential sur-
others (such as the amide I and II peaks) is face of deformational modes while flattening the
enhanced slightly. potential surface of stretching modes.29 In the
3. At 35–37 °C, a phase-transition region, in- deconvolution, peak positions and peak shapes
dicated by a sharp increase in the absor- (20% Lorentzian and 80% Gaussian) were taken
bance of all the PNIPA-related peaks. as constant parameters, whereas the intensity
4. Above 37 °C, a posttransition region, man- and width of the amide I and II subbands were
ifested by a very gradual increase of all regarded as free variables.
peaks. Figure 6 presents the deconvoluted IR spectra
of amide I and amide II peaks of a 2.5% PNIPA
It is well known that the solubility of PNIPA in solution at 25 and 40 °C. A difference in the peak
water is due to the hydrophilic amide groups that shapes is clearly observed. A detailed, quantita-
possess the capability of hydrogen bonding either tive examination of the temperature-dependent
with the surrounding water molecules or with relative contribution of each of the amide II com-
other amide groups. It is further known that, for ponents is presented in Figure 7. The figure re-
polyamides, the band contour of the amide I and veals that changes in the relative intensity of the
amide II peaks are conformationally sensitive.21(a) subbands exist at temperatures lower than the
In agreement with the treatment of Lin et al.,23 phase-transition temperature and may provide a
the amide I peak (mostly CAO stretch) in the hint about the collapse scenario. Four different
subtraction spectra was deconvoluted into three regions can be observed:
subbands: an intramolecular hydrogen-bonded
subband at 1630 cm⫺1, an intermolecular hydro- 1. Below 31 °C, increasing temperature
gen-bonded subband at 1620 cm⫺1, and a free causes a moderate decrease in the intermo-
form of non-hydrogen-bonded subband at 1643 lecular hydrogen bonds and a moderate in-
cm⫺1. Likewise, the amide II band (mostly NOH crease in the free amide population. The
deformation) was deconvoluted into an intramo- intramolecular hydrogen-bond population
lecular hydrogen-bonded subband at 1551 cm⫺1, hardly changes.
an intermolecular hydrogen-bonded subband at 2. At 32–35 °C, there is an abrupt decrease in
1565 cm⫺1, and a free form of non-hydrogen- the intermolecular hydrogen-bond popula-
ATR/FTIR STUDIES OF PNIPA 1671

Figure 7. Changes in the three subbands of the amide II peak as a function of


temperature for a 2.5% PNIPA solution: (Œ) intermolecular hydrogen-bonding subband,
(F) free amide groups, and (䊐) intramolecular hydrogen-bonding subband).

tion and an abrupt increase in the intramo- fact that the relative integral of the intramolecu-
lecular hydrogen bond and the free amide lar hydrogen-bonding subband increased four
populations. This region is characterized times from about 7% to about 28% implies that at
also by an abrupt increase in the width of low temperatures the percentage of amide groups
the free amide subband. engaged in intramolecular hydrogen bonding does
3. At 36 –39 °C, there is an increase in the not exceed 25% of the total number of amide
relative amount of the intermolecular hy- groups.
drogen-bond population and a decrease in
the integrated absorbance and width of the
free amide subband. Phase-Separation Effect as Manifested by the
4. Above 40 °C, there is a moderate increase Water-Related IR Peaks
in the relative population of the free amide, In the mid-IR range, water has three bands: an
coupled with a moderate decrease in the HOH bending mode at 1640 cm⫺1, a wide combi-
intramolecular hydrogen-bond and inter- nation band (1800 –2500 cm⫺1) peaking at 2125
molecular hydrogen-bond populations. cm⫺1, and a strong OOH stretch band at 3000 –
3700 cm⫺1. The IR spectrum of water is known to
The same trends were observed in an analysis change with increasing temperature because of
of the temperature-dependence graphs of the in- the reduced extent of hydrogen bonding. The
tegrated intensity of the amide I subbands, al- main changes are the shifting of the OOH enve-
though amide I graphs were found to be less lope toward higher frequencies and the narrowing
smooth than the amide II graphs, probably the of the HOH bending peak, accompanied by a
result of some remaining overlap with the HOH small change in this peak toward lower frequen-
bending. The same four regions were also found in cies. It was already mentioned that the 2125-
aqueous solutions containing PNIPA at other con- cm⫺1 peak was used for internal calibration to get
centrations. an accurate subtraction of the water spectra from
Although the relative integrated absorbance that of the PNIPA solution. On the basis of this
should not be taken as representing the relative analysis, an apparent sharp decrease in the water
quantities of the different populations (because of signal was observed with the phase transition.
a possible difference in the transition moment of This complementary section discusses changes in
the three populations), it is still possible to obtain the temperature-induced behavior of water due to
some quantitative information. For example, the interaction with solvated PNIPA.
1672 RAMON ET AL.

in pure water decreased as a function of temper-


ature because of the breaking of hydrogen bonds.
Such monotonic behavior was observed also in the
PNIPA solution, although in the latter case, a
sharp decrease at the phase-transition tempera-
ture was observed. Furthermore, a change in the
slope of the graph, to a more pronounced de-
crease, was observed at temperatures above the
transition point.
In accordance with the procedure of Sammon
et al.,30 the OOH stretch band of pure water and
water containing PNIPA was deconvoluted into
four subbands, reflecting four distinct types of
water environments, that differed by the degree
of hydrogen bonding (Fig. 9). Although this clas-
sification is somehow artificial, it does follow ex-
perimental evidence showing that strong hydro-
gen bonding skews the 3000 –3700-cm⫺1 envelope
toward lower wavenumbers, whereas a lesser ex-
tent of hydrogen bonding skews that envelope
toward higher wavenumbers. In the deconvolu-
tion, the subbands were (a) 3223 cm⫺1, strong
hydrogen bonding; (b) 3395 cm⫺1, medium-
strength hydrogen bond; (c) 3525 cm⫺1, weak hy-
drogen bond; and (d) 3612 cm⫺1, free water. In a
fashion similar to the procedure taken during the
deconvolution of the amide I and amide II peaks,
the peak positions and peak shapes (20% Lorent-
zian and 80% Gaussian) were taken as constant
parameters, whereas the intensity and width of
the subbands were regarded as free variables. To
verify that the NH stretch bands of the PNIPA
and its amide II overtone did not alter the anal-
Figure 8. Integrated intensity of water-related peaks ysis, we took care to include these peaks as well,
in (E,‚) pure water and (F, Œ) a 2.5% PNIPA solution: but their contribution was found to be negligible,
(A) HOH bending mode (1635 cm⫺1) and (B) OOH
compared with the water signal.
stretch envelope (3000 –3700 cm⫺1).
The relative areas of the OH stretch subbands
of water and the relative areas of the OH stretch
subbands of water containing 2.5% PNIPA are
The integrated absorbance of the HOH bending presented, as a function of temperature, in Figure
mode (1640 cm⫺1) of water and water containing 10. As shown in the figure, at a temperature of 25
dissolved PNIPA is presented in Figure 8(A) as a °C, the area of subband a was close to 55% of the
function of temperature. We calculated the area total area of the peak, whereas the areas of sub-
of the HOH peak of the PNIPA solution by sub- bands b– d were 33, 10, and 2%, respectively. Wa-
tracting the area of the amide signals from the ter in the PNIPA solution was found to differ from
overall area of the peak. As shown in the figure, pure water by a lower extent of strong hydrogen
the integrated intensity of pure water increased bonds and a higher extent of medium-strength
slightly with temperature. In contrast, the inte- hydrogen bonds. In contrast, hardly any differ-
grated intensity of this peak in the PNIPA solu- ence could be observed in the extent of weak hy-
tion exhibited a maximum at the phase-transition drogen bonds or free water. The lower amount of
temperature and decreased at temperatures strong hydrogen-bonding subbands found for the
higher than the LCST. As shown in Figure 8(B), PNIPA solution correlates well with the claim
the area of the coupled intramolecular and inter- that PNIPA behaves like a structure breaker, be-
molecular OOH stretch band (3000 –3700 cm⫺1) cause the water must reorient around the nonpo-
ATR/FTIR STUDIES OF PNIPA 1673

Figure 9. Deconvoluted spectrum of the OH stretch envelope of water, showing the


four types of populations.

lar regions of the solute.31 At low temperatures, (1–3.5%). The phase-transition temperature was
the differences between pure water and PNIPA found to be the same, within an experimental
solutions can hardly be observed by an inspection error of 1 °C, for all solutions. Qualitatively, the
of the crude data (Fig. 3) because of the small spectral changes were similar, regardless of poly-
concentration of solute. However, these differ- mer concentration. These changes included an
ences are easy to observe if one inspects the sub- increase in the PNIPA IR signals, a decrease in
traction spectra (Fig. 4), where a negative peak is water-related signals, and a similar behavior for
observed at the locus of the strong hydrogen- deconvoluted amide peaks. An interesting point
bonding subband. was the increase of the PNIPA peaks with the
As the temperature of the liquid is increased, a phase transition. Below the transition tempera-
decrease in the 3223-cm⫺1 subband of pure water ture, the integrated absorbance of the PNIPA
could be observed, together with an increase in peaks increased, more or less, linearly with the
the 3395-cm⫺1 subband. Slight changes were concentration, as could be expected. In contrast,
found also in the other two subbands (the 3525- the integrated absorbance at higher tempera-
cm⫺1 subband decreased, and the 3612-cm⫺1 sub- tures was almost constant, regardless of the
band increased); however, these changes were PNIPA concentration. This constant absorbance
smaller than the deconvolution error. A similar as a function of concentration occurred for all the
behavior was found also for water containing PNIPA-related peaks.
PNIPA. At the phase-transition point, however,
an abrupt decrease in the relative amounts of the
strong hydrogen-bonding subband and weak hy-
DISCUSSION
drogen-bonding subband coupled with an abrupt
increase in the relative amount of medium-
Reduction in Water Concentration near the ATR
strength hydrogen-bonding subband was ob-
Crystal
served. At higher temperatures, a subtle increase
in the slope of the temperature-dependent rela- As described in the previous sections, the phase
tive areas could be observed. transition was characterized by a sharp increase
in the ATR signal of all the PNIPA-related peaks
(Fig. 4). This increase at the phase-transition
Effect of PNIPA Concentration on Phase-Transition
temperature was coupled with a decrease in all
Phenomena as Measured by ATR
the water-related peaks: the HOH bending at
As already mentioned, the measurements were 1640 cm⫺1 [Fig. 8(A)], the OH stretch envelope at
performed at various concentrations of PNIPA 3000 –3700 cm⫺1 [Fig. 8(B)], and the wide combi-
1674 RAMON ET AL.

tor (for a given concentration) suggests that the


observed increase in the signal was mostly due to
an increase in the average concentration of the
polymer close to the crystal surface. This sugges-
tion is supported by the decrease in all the water-
related signals at the phase-transition tempera-
ture. Furthermore, the fact that the area of the
PNIPA peaks was, above the transition tempera-
ture, independent of polymer concentration in the
solution may suggest that a film with a constant
PNIPA concentration is formed with the phase
transition. If indeed this is the case, the average
PNIPA concentration near the crystal (at a dis-
tance of a few micrometers) should be 7–10%,
based on calibration according to the IR signals of
the PNIPA bands as measured below LCST. This
value is further supported by the water peak
data.

Phase-Transition Scenario
The temperature-induced changes in the relative
areas of the three subbands in the amide II peak
(Fig. 7) imply a four-stage process of phase sepa-
ration in the PNIPA solution, which will be
termed the early (A), intermediate (B), aggrega-
tion (transition temperature), relaxation (C), and
postrelaxation (D) stages.
The early stage, below 31 °C, is characterized
by the domination of the water–PNIPA intermo-
lecular hydrogen bonding, which contributes to
Figure 10. Relative area of the four subbands in the approximately 60% of the overall amide II signal
OH stretch envelope for pure water (empty symbols) measured at this stage. As the temperature is
and a 2.5% PNIPA solution (filled symbols): (䊐,■) raised, the relative contribution of this subband is
strong hydrogen-bonding subband (3223 cm⫺1), (‚,Œ) gradually reduced, whereas that of the subband
medium-strength hydrogen-bonding subband (3395 corresponding to non-hydrogen-bonded amides
cm⫺1), (〫,⽧) weak hydrogen-bonding subband (3525 (free amides) is increased. Because of the in-
cm⫺1), and (E,F) free-water subband (3612 cm⫺1). creased number of free amides, the movement of
the side chains becomes less restricted, a factor
that may have an influence on the flexibility of
nation band at 2125 cm⫺1 (see the description on the backbone of the polymer. The required in-
the subtraction procedure). The aggregation of crease in entropy during this stage could result
molecules is known to have strong effect on band from a decrease in the number of ordered water
intensity. However, in this case it is expected that molecules due to the breaking of the thin shell of
each band will respond to aggregation in a differ- ordered molecules around the polymer. However,
ent manner, according to the change in its in- it is possible that during this pretransition stage,
duced dipole moment. In particular, it could be the required entropy comes from structural
expected that the extent of the increase would changes in the backbone of this relatively flexi-
differ significantly between NH-related peaks ble32 polymer, as water, considered to be a good
and COH-related peaks, or between stretch solvent at low temperatures, becomes a ␪ solvent
modes and deformation modes. The fact that the at ⬃32 °C. Indeed, both the radius of gyration and
area of all PNIPA peaks, including those related the hydrodynamic radius are known to decrease
to the hydrophobic groups, were found to increase moderately at this stage,33 approaching maximal
at the transition approximately by the same fac- entropy at the ␪ temperature.
ATR/FTIR STUDIES OF PNIPA 1675

The intermediate stage (31–35 °C) is charac- lapsed globules contain less water22,34and expose
terized by a sharp decrease in the relative inten- their isopropyl groups to the solvent, which be-
sity of the intermolecular hydrogen-bonding sub- comes less polar in the vicinity of the polymer, as
band. This decrease is coupled with a sharp in- demonstrated by Winnik.35 This hydrophobically
crease in the amide–amide hydrogen-bond driven aggregation mechanism does not contra-
population in the lower temperatures of this stage dict the mechanism suggested by us but rather is
and an increase in the population of the free complementary to the one presented here.
amide groups at the higher temperatures of this Above the transition temperature and up to 40
stage. The phase separation, as indicated by DSC °C, a decrease in the amount of free amide groups,
measurements and visual inspection, occurred coupled with an increase in the intermolecular
quite abruptly at the end of this stage, that is, at component (i.e., water–amide hydrogen bonding),
35 °C. The behavior below the phase-separation could be observed. It is believed that at this re-
temperature suggests that the phase-separation laxation stage hydrogen bonds are formed be-
process is triggered by the formation of free amide tween water molecules that were trapped inside
groups. Once a critical number of side chains are the polymer globules and amide groups that were
free to change their conformation, they begin to free. Above the transition temperature, the rela-
form amide–amide hydrogen bonds. It is likely tive amount of water molecules belonging to the
that at this stage the amide–amide bonds are weak hydrogen-bonding population is decreased,
intramolecular because, otherwise, one could whereas that of the medium-strength hydrogen-
have expected this scenario to be concentration- bonding population is increased (Fig. 10). It is not
dependent. The structural changes involved in clear whether this difference is due to water
the formation of the intramolecular hydrogen trapped within the globules or water molecules at
bonds trigger a very fast decrease in the intermo- the interface between bulk water and polymer
lecular water–amide bonds, thus further increas- aggregates. A true equilibrium is achieved only
ing the amount of free amides. At this point, the above 40 °C, as demonstrated in Figure 7. Hence,
polymer side chains are expected to be very flex- it can be said that although phase separation in
ible, thereby able to form hydrophobic interac- PNIPA is well defined and abrupt, the scenario of
tions between their isopropyl groups and/or their this process can be quite complex and occurs over
backbone. Because in PNIPA the hydrophobic iso- a wide range of temperatures.
propyl groups are located adjacent to the amide It is interesting to compare these findings with
groups, one could assume that the formation of those of other groups who used different tech-
the intramolecular hydrogen bonding at 33 °C niques. On the basis of an analysis of the temper-
should be coupled with the formation of hydro- ature dependence of the ratio between the radius
phobic interactions, however, no change in the of gyration and the hydrodynamic radius, as ob-
intensity or vibration frequency of methyl or tained from light scattering experiments, it has
methylene groups could be observed at tempera- been claimed that, before aggregation, the chains
tures below the transition temperature. go through a four-stage scenario where the chains
Phase separation (i.e., aggregation of the col- contract, crumple, rearrange, and collapse.33,34
lapsed chains) occurred at a point where the
amount of free amide groups was at its maximum.
This point was characterized also by a shift in the CONCLUSIONS
vibrational band of the hydrophobic groups, prob-
ably reflecting the formation of intermolecular This article has demonstrated that the ATR/FTIR
hydrophobic interactions. As can be deduced from technique can provide detailed information on the
analyzing the water-related peaks, this aggrega- process of temperature-induced phase separation
tion stage is characterized also by the repulsion of in aqueous solutions of hydrophobic polymers,
water from the vicinity of the chains and the such as PNIPA. Based on the ATR/FTIR results,
formation of polymer-rich regions on the surface the following scenario for the temperature-
of the ZnSe crystal. This scenario is in line with induced collapse of PNIPA is proposed.
claims made by other groups, according to which At low temperatures, the polymer is solvated in
the PNIPA–water system is thermodynamically such a manner that the hydrophilic groups form
stable at low temperatures, and it becomes unsta- relatively strong hydrogen bonds with water. As
ble at the onset of the phase-separation stage.34 It the temperature is raised, more and more hydro-
has been suggested that at this stage the col- gen bonds between amide groups and water are
1676 RAMON ET AL.

broken. On average, the free amides do not form 7. Shibayama, M.; Tanaka, T. J Adv Polym Sci 1993,
new intramolecular hydrogen bonds as long as 109, 1– 62.
their number is low. Once a critical number of 8. (a) Cole, C.-A; Schreiner, S. M.; Priests, J. H.; Mu-
bonds are broken, intramolecular hydrogen bonds nji, N.; Hoffman, A. S. Am Chem Soc Symp Ser
are formed, at the expense of the free amide group 1987, 350, 245–254; (b) Priest, J. H.; Murrey, S. L.;
population and the intermolecular hydrogen-bond Nelson, R. J.; Hoffman, A. S. Am Chem Soc Symp
Ser 1987, 350, 255–264; (c) Matsukata, M.; Aoki,
population. As a result, the polymer begins to
T.; Sanui, K.; Ogata, N.; Kikuchi, A.; Sakurai, Y.;
transform itself to a more compact conformation.
Okano, T. Bioconjugate Chemistry 1996, 7, 96 –101.
This promotes the breaking of more intermolecu-
9. Yamamoto, I.; Iwasaki, K.; Hirotsu, S. J Phys Soc
lar hydrogen bonds, which eventually causes the Jpn 1989, 58, 210 –215.
CGT of the chain. The exposed hydrophobic moi- 10. (a) Walker, J. A.; Vause, C. A. Sci Am 1987, 253, 98;
eties induce aggregation due to hydrophobic in- (b) Hirotsu, S. J Phys Soc Jpn 1987, 56, 233–242.
teractions. The collapsed, aggregated polymer 11. (a) Tanford, C. The Hydrophobic Effect, 2nd ed.;
tends to adhere to the ATR crystal, so an increase Wiley: New York, 1973; Chapters 1–3; (b) Ulbrich,
in the intensity of all polymer-related peaks is K.; Kopecek, J. J Polym Sci Polym Symp 1979, 66,
observed, together with a decrease in the inten- 209; (c) Ben-Naim, A. Hydrophobic Interactions;
sity of all water-related peaks. On the basis of the Plenum: New York, 1980.
decrease in the free amide signal and the increase 12. (a) Heskins, M.; Guillet, J. E. J Macromol Sci Chem
in the signal of the intermolecular hydrogen 1968, 2, 1441–1455; (b) Schild, H. G.; Tirell, D. A. J
bonds above the phase-separation temperature, it Phys Chem 1990, 94, 4352– 4356.
is believed that, with phase separation, water 13. (a) Ricka, J.; Meenes, M.; Nyffenegger, R.; Binkert,
molecules that were entrapped within the col- T. Phys Rev Lett 1990, 65, 657– 660; (b) Meewes,
lapsed areas form hydrogen bonds with free M.; Ricka, J.; Le Silva, M.; Nyffeneger, R.; Binkert,
amide moieties. T. Macromolecules 1991, 24, 5811–5816.
14. (a) Inomata, H.; Goto, G.; Saito, S. Macromolecules
The detailed role of water in the phase-transi-
1990, 23, 4887– 4888; (b) Tiktopulo, E. I.; Bych-
tion process of PNIPA solutions has been revealed
kova, V. E.; Ricka, J.; Ptitsyn, O. B. Macromole-
(to our knowledge) for the first time on a molecu-
cules 1994, 27, 2879 –2882; (c) Boutris, C.; Chatzi,
lar level by the CCATR/FTIR technique. At E. G.; Kiparissides, C. Polymer 1997, 38, 2567–
present, this technique is used to study the effect 2570.
of low molecular weight solutes (salts) on that 15. Fujishige, S.; Kubota, K.; Ando, I. J Phys Chem
process. A forthcoming article on that subject is 1989, 93, 3311.
now under preparation. 16. (a) Winnik, F. M. Macromolecules 1989, 22, 734 –
742; (b) Schild, H. G.; Tirrell, D. A. Langmuir 1991,
This work was performed under contract number 7, 665– 671; (c) Schild, H. G.; Tirrell, D. A. Macro-
205-97 of the Israel Science Foundation. The authors molecules 1992, 25, 4553– 4558; (d) Armentrout,
thank Professor Rolfe Herber (Hebrew University) for R. S.; Hu, Y.; McCormick, C. L. Polymer Preprints
lending the circle cell and for a critical reading of the 1996, 37, 659 – 660.
manuscript. 17. (a) Tokuhiro, T.; Amiya, T.; Mamada, A.; Tamaka,
T. Macromolecules 1991, 24, 2936 –2943; (b) Zeng,
F. Z.; Feng, H. Polymer 1997, 38, 5539 –5544.
REFERENCES AND NOTES 18. Scarpa, J. J.; Mueller, D. D.; Klotz, I. M. J Am
Chem Soc 1967, 89, 6024 – 6030.
1. Piculell, L.; Nilsson, S. Prog Colloid Polym Sci
1990, 82, 198 –210. 19. Snyder, W. D.; Klotz, I. M. J Am Chem Soc 1975,
2. Stockmayer, W. H. Makromol Chem 1960, 35, 54. 89, 4999 –5003.
3. (a) Grosberg, A. Y.; Kuznetsov, D. V. Macromole- 20. (a) Marechal, Y. J Mol Struct 1994, 322, 105–111;
cules 1993, 26, 4249 – 4251; (b) Yamakawa, H. Mac- (b) Rossi, A. V.; Dawanzo, C. V.; Tubino, M. J Braz
romolecules 1993, 26, 5061–5066; (c) Park, I. H.; Chem Soc 1996, 7, 403– 410.
Wang, Q.; Chu, B. Macromolecules 1987, 20, 1965– 21. (a) Skrovanek, D. J.; Howe, S. E.; Painter, P. C.;
1975; (d) Chu, B.; Park, I. H.; Wang, Q. W.; Wu, C. Coleman, M. M. Macromolecules 1986, 19, 1676 –
Macromolecules 1987, 20, 2833–2840. 1683; (b) Skrovanek, D. J.; Painter, P. C.; Coleman,
4. Tanaka, T. Polymer 1979, 20, 1404 –1412. M. M. Macromolecules 1986, 19, 699 –705; (c)
5. Kubota, Ki.; Fujishige, S.; Ando, I. Polym J 1990, Durrani, C. M.; Prystupa, D. A.; Donald, A. M.
22, 15–20. Macromolecules 1993, 26, 981–987; (d) Harris,
6. For a comprehensive review on PNIPA, see Schild, P. I.; Chapman, D. Trends Int Biol Sci 1992, 17,
H. G. Prog Polym Sci 1992, 17, 63–249. 328 –333.
ATR/FTIR STUDIES OF PNIPA 1677

22. Shibayama, M.; Marimoto, M.; Nomura, S. Macro- 29. Vinogradov, S. N.; Linnel, R. H. Hydrogen Bond-
molecules 1994, 27, 5060 –5055. ing; Van Nostrand Reinhold: New York, 1971; pp
23. Lin, S.-Y.; Chen, K.-S.; Run-Chu, L. Polymer 1999, 47– 82.
40, 2619 –2624. 30. Sammon, C.; Mura, C.; Yarwood, J.; Everall, N.;
24. Otake, K.; Inomata, H.; Konno, M.; Saito, S. Mac- Swart, R.; Hodge, D. J Phys Chem B 1998, 102,
romolecules 1990, 23, 283–289. 3402–3411.
25. Hirokawa, Y.; Tanaka, T. J Chem Phys 1984, 81, 31. Uberreiter, K. Colloid Polym Sci 1982, 260, 37– 45.
6379 – 6380. 32. Kubota, K.; Fujishige S.; Ando, I. Polym J 1990, 22,
26. Libnau, F. O.; Toft, J.; Ehristy, A. A.; Kvalheim, 15–20.
O. M. J Am Chem Soc 1994, 116, 8311– 8316. 33. Wu, C.; Zhou, S. Macromolecules 1995, 28, 8381–
27. Rahmelow, K.; Hubner, W. Appl Spectrosc 1997, 8387.
51, 160 –170. 34. Wu, C.; Zhou, S. Macromolecules 1995, 28, 5388 –
28. Elliot, A. Infra-Red Spectra and Structure of Or- 5390
ganic Long-Chain Polymers; St Martin’s Press: 35. Winnik, F. M. Macromolecules 1990, 23, 233–
New York, 1969; pp 86 – 87. 242.

Anda mungkin juga menyukai