Você está na página 1de 19

Acrylic Acid Polymerization Kinetics

S. S. CUTIE, 1 P. B. SMITH, 1 D. E. HENTON, 2 T. L. STAPLES, 2 C. POWELL 2


1

Analytical Sciences

Superabsorbent Products R&D, The Dow Chemical Company, Midland, Michigan 48674

Received 10 May 1996; revised 27 March 1997; accepted 31 March 1997

ABSTRACT: The kinetics of the isothermal polymerization of acrylic acid were deter-

mined utilizing 1H-NMR spectroscopy. The polymerization rate was observed to depend
approximately on the 32 power of monomer and the 12 power of sodium persulfate concentration. This is consistent with a model in which the rate of initiation is itself dependent
on the monomer concentration. The polymerization rate was also observed to have
a strong dependence on percent neutralization, decreasing with increasing level of
neutralization up to 75 to 100% neutralization, and then increasing again. The activation energy for the rate of polymerization was between 9 and 13 kcal/mol except for
100% neutralized acrylic acid, which had an activation energy of 18 kcal/mol. These
data suggest that a transition in mechanism occurred at 100% neutralization. Increasing the ionic strength by the addition of sodium chloride also increased the rate. The
dependence of the molecular weight on the above variables was also quantified for use
in the model. It decreased with increasing conversion, decreasing ionic strength and
increasing initiator. q 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35: 2029
2047, 1997

Keywords: poly(acrylic acid) gel; polymerization kinetics; nuclear magnetic resonance;


crosslinking; superabsorbent

INTRODUCTION
Improving the performance of superabsorbent
polymers depends, to a considerable extent, on
controlling the polymerization. Obtaining high
backbone molecular weight, uniform crosslinker
incorporation, low residual monomer levels, and
high production rates, all at the same time, requires a detailed knowledge of the polymerization
kinetics.
Several groups previously investigated the
polymerization of acrylic acid. Ito et al. reported a
large pH dependence of the rate of polymerization
with a minimum near pH 7.1 Kabanov et al. also
observed that the rate of polymerization of acrylic
acid decreased as the pH was raised from 1 to 6,

Correspondence to: P. B. Smith


Journal of Polymer Science: Part B: Polymer Physics, Vol. 35, 20292047 (1997)
q 1997 John Wiley & Sons, Inc.
CCC 0887-6266/97/132029-19

then increased again until pH 12.2,3 Above pH 12


they noted that the rate again decreased. The authors proposed that the polymerization rate was
enhanced by the presence of ion pairs whose concentration increased with increasing pH. These
authors also noted that the stereochemistry of the
polymer seemed to be controlled by this ion pair
formation and that the addition of high concentrations of salts such as NaCl and CH3COONa increased both the polymerization rate and the molecular weight of the linear polymer.
Manickam et al. also observed an enhancement
of the polymerization rate of acrylic acid with the
addition of salt, but only in alkaline medium.4,5
The salt effect was not observed at low pH values.
Laborie, Chapiro, and Dulieu, and Galperina
et al. suggested that the polymerization in aqueous medium was accelerated by a matrix effect
in which monomers associate prior to polymerization into H-bonded aggregates.6 9 In contrast,
2029

/ 8Q31$$5007

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

2030

ET AL.
CUTIE

hydrocarbon and chlorinated solvents, which do


not support these H-bonded aggregates, lead to a
markedly reduced rate of polymerization and result in lower molecular weight. This matrix effect was found to persist to fairly high dilutions
and was proposed as the origin of stereochemical
control of the resulting polymer.
Acrylamide is another water-soluble monomer
that yields very high molecular weight polymers.
The rate of polymerization of neutral acrylamide
was shown by Riggs and Rodriguez to have a
strong dependence on the cation of added salts,
potassium causing about twice the increase in
rate as sodium.10 Lenka et al. modeled the polymerization of acrylamide by peroxydiphosphate
and showed that its polymerization did not require a matrix effect to fit the data.11 Prabha and
Nandi interpreted the selectivity of the initiator,
ferric dipivaloylmethide, toward a given vinyl
monomer in its polymerization and spectroscopic
data, as evidence of complex formation between
the initiator and the monomer.12
This article describes the determination of the
isothermal polymerization rate constants for
acrylic acid as a function of pH, monomer concentration, and temperature. A model, which incorporates these effects, is presented.

EXPERIMENTAL
A 1H-NMR technique was used to monitor the polymerization rates of acrylic acid in situ.13 This
method made it possible to obtain isothermal data
during exothermic conditions of up to 10% conversion/minute. Small feed batches of 50100 g were
prepared in glass bottles, as described below, with
the appropriate initiator level, monomer content,
percent neutralization, and additives.
Typical Feed Batch Preparation Procedure for
31.9% Solids, 65% Neutralization, 1600 ppm
Sodium Persulfate Based on Acrylic Acid (BOAA)
To 28.18 g of glacial acrylic acid (monomer grade,
200 ppm MEHQ) was added 69.925 g of the persulfate master batch solution.
Master Batch of Sodium Persulfate
698.8 g DI water; 0.450 g sodium persulfate
(98/%, Aldrich).
This mixture was neutralized by the slow addi-

/ 8Q31$$5007

07-28-97 20:31:22

Figure 1. The 1H-NMR spectra of the polymerization


of acrylic acid as a function of time.

tion, with stirring, of 13.456 g of granular sodium


carbonate (Fisher Certified A.C.S.).
NMR Kinetic Analysis
The appropriate diameter NMR tube (depending
on heat transfer requirements) was loaded via a
glass pipette and deoxygenated with nitrogen in
a microdevice described elsewhere.13 The tubes
were deoxygenated for a minimum of 30 min.
After the deoxygenation, the cap was sealed and
the sample was sonicated (2030 s) to remove
the nitrogen gas from the saturated liquid so that
bubbles did not form in the sample as it polymerized. (Bubbles give rise to severe line broadening
in the 1H-NMR spectra, dramatically degrading
their quality.) The sample was then placed in the
temperature-equilibrated, nitrogen-filled probe of
the NMR spectrometer and data were acquired.
The time of data accumulation for each spectrum
was determined by the number of scans and the
delay time between scans. A typical series of 1HNMR spectra, from low to high conversion are
given in Figure 1. Of the different NMR techniques, 1H-was preferred over 13C because of the
greater sensitivity, faster acquisition rates, and
the ability to use smaller diameter tubes, which
helped maintain isothermal conditions. The concentration of polymer was determined from the
areas of the peaks at 1.52.5 ppm and that of the
monomer from the area of the peaks at 5.56.5
ppm. The fractional conversion was calculated as
the level of polymer area divided by the sum of
the areas of monomer and polymer. The rate of

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

conversion was typically calculated from the initial slopes of these curves. The rate of polymerization (Rp ) was obtained by multiplying the rate of
conversion by the initial monomer concentration.
The temperature of the probe was set to within
{17C using ethylene glycol as an internal reference; the 1H-NMR chemical shift difference between the methylene and hydroxyl protons of ethylene glycol gives an accurate measure of temperature. The polymerization is very exothermic, and
the sample temperature was found to increase
above the NMR set point when the rate of reaction
exceeded about 2% per minute for samples at 30%
solids or higher in 5 mm NMR tubes. This was
partly due to the limitations of heat transfer from
within the sample to the NMR tube but mostly to
the ability to transfer the heat from the NMR tube
within the NMR probe. The temperature within
the NMR probe was maintained by temperature
equilibrated nitrogen gas that was forced around
the sample. The heat transfer characteristics of
the gas were not sufficient to manage the large
exotherms in the sample.
These observations indicated that, under some
conditions, even the NMR tube reactor did not
allow adequate removal of the heat generated by
the polymerization to maintain isothermal conditions. A special NMR tube configuration was used
for samples that were expected to polymerize at
rates faster than about 2% per minute. This apparatus was composed of a 2 mm i.d. NMR tube
coaxially inserted into a 5 mm o.d. NMR tube. The
sample was placed into the 2 mm tube and the
volume between the 2 and 5 mm tubes was filled
with D2O. The D2O acted both as a heat transfer
fluid and as an internal lock for the NMR. This
apparatus provided better heat transfer and also
minimized the sample size so that less heat was
generated.
The 1H-NMR spectra were obtained at 299
MHz using a Bruker AC-300 NMR spectrometer,
model number HO2129-ECL-24, S/N 0898. The
data acquisition parameters utilized were as follows: pulse width 107, delay time 10 s, size
16 K, accumulation time 1.36 s, sweep width
6 KHz, apodisation exponential, 0.5 Hz
broadening.
Molecular Weight Analysis
The size-exclusion chromatography (SEC) system
consisted of two TosoHaas TSK columns GPWXL ,
and one TSK column G 1000 PW.15 TSK gel
GPWXL columns are packed with G 2500, G3000

/ 8Q31$$5007

07-28-97 20:31:22

2031

and G6000 PWXL resins. The eluent used was


0.3N NaCl with 0.03M Na2HPO4 . The pH was
maintained at 6.8 with 3% phosphoric acid. The
solvent was filtered through a Micron Separations
Inc. 0.1 mm Nylon filter available from Fisher Scientific, Midland, MI. Sodium azide (400 ppm) was
added to inhibit bacteria formation. The pump
was a Waters 590 with excellent flow reproducibility. The flow was kept at 1 mL/min. The detector
was a Waters refractive index model 410. The columns and detector cells were kept at 357C. The
injection volume was 200 mL. Samples were filtered through a 0.45 mm Nylon Acrodiscy 13 mm,
available from Fisher Scientific, Midland, MI.
The apparent molecular weight of the samples
was determined by comparison of the chromatographic behavior of the sample to that of standards with known molecular weights. Polyethylene oxide narrow distribution standards ranging
in molecular weight from 1470 to 1,390,000 were
used for calibration. The molecular weight data
above about 1,400,000 Daltons should be used
with caution due to the lack of standards.
To measure the molecular weight distribution
of the crosslinked polymer, the crosslinks needed
to be hydrolyzed to produce a soluble polymer.
The crosslinked polymer (0.04 g) was added to
0.1N NaOH (10 mL) in water (1/250 ratio by
weight), MEHQ (0.04 g) was added, and placed
in an oven at 757C for 72 h to undergo hydrolysis.
After completion of the hydrolysis the sample was
1
diluted 10
with SEC eluent and injected into the
system. This procedure led to the hydrolysis of
crosslinks composed of ester bonds but not the
carbon carbon crosslinks formed by chain
transfer.

RESULTS
Polymerization Rate

Effect of Monomer and Initiator Concentrations


The isothermal polymerization of acrylic acid in
water was monitored as a function of concentration, temperature, degree of neutralization and
initiator level to define a kinetic model and obtain
accurate values for the rate constants. The dependence of the polymerization rate on initial monomer concentration was determined at 557C by
measuring the rates of polymerization for feed
batches containing acrylic acid concentrations
ranging from 37.2% (4.56M) to 9.3% (1.17M) at

polpa

W: Poly Physics

9605007

2032

ET AL.
CUTIE

Figure 2. The rate of polymerization of 65% neutralized acrylic acid at 557C as a


function of acrylic acid concentration.

constant total persulfate concentration. The


monomer concentration in wt % was based on the
weight of the 65% neutralized monomer. A plot of
the conversion as a function of time for three such
runs is given in Figure 2. It shows that the rate
of conversion increases with increasing initial
monomer concentration. Radical polymerizations
generally possess a propagation rate that is proportional to kp[M][R], where kp is the propagation rate constant. This step is almost always first
order in both the monomer [M] and the growing
radical [R]. Over any significant time period, radical concentrations will generally be constant or
only slowly changing; this gives rise to the socalled steady state, and in such cases the conversion with time curve will be independent of monomer concentration.
More complex rate expressions arise when the
steady-state radical concentration itself depends
on the monomer concentration. This leads to other
than first order kinetics, which has been found by
others to be the case for acrylic acid. Ito et al.
cited a rate law with second-order monomer dependence.1 Manikam et al.5 observed a 32 power
dependence on monomer, as did Kabanov et al.3
Interestingly, the latter paper showed a simple
first order dependence on monomer for methacrylic acid.
If the undissociated initiator interacts with the
monomer, either due to a redox couple, induced
decomposition, or because the initiator fragments
are trapped in a solvent cage, the rate of initiation

/ 8Q31$$5007

07-28-97 20:31:22

will depend on the monomer concentration [M].


Because bimolecular termination typically controls the steady state concentration of radicals
such that [R] depends on the square root of the
rate of initiation, it follows that the overall dependence of propagation would show a 32 dependence,
one power comes from the normal monomer dependence, and the additional half power from the
rate of initiation.
q

[R ] fkd[I][M]

(1)

where f has been introduced as an efficiency factor and kd[I] is the normal first-order rate of decomposition of initiator. It should be noted that
this behavior may be particularly dependent on
the specific initiator used.
The logarithm of the initial polymerization rate
for these experiments is plotted as a function of
initial monomer concentration in Figure 3. The
slope of the curve gives a 1.73 dependence on
monomer concentration that is approximately
consistent with the 32 dependence discussed above.
A line with a slope of 1.5 is shown in Figure 3 for
comparison. The discussion of the polymerization
model will elaborate on this topic further.
The dependence of the polymerization rate on
the persulfate initiator concentration was determined at 557C, 65% neutralization, and 33.8% solids, as shown in Figure 4. The rate of polymerization increases, as expected, as the persulfate initi-

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

2033

the O2 is consumed by the radicals, no polymerization takes place.13 This side reaction is believed
to be the cause of the behavior of the plot.

Degree of Neutralization and Temperature

Figure 3. Effect of monomer concentration on the initial rate of acrylic acid polymerization at 557C and 65%
neutralization.

ator level increases because more chains are


initiated. A plot of the initial rate of polymerization vs. persulfate concentration to the 12 power
(Fig. 5) gives a straight line, verifying the half
power dependence for this system proposed in the
equation above.
The plot of the initial rate of polymerization vs.
persulfate concentration given in Figure 5, does
not go through the origin. There are components
in the acrylic acid that scavenge the radicals initially. Commercial acrylic acid has approximately
200 ppm MEHQ, which works synergistically
with oxygen to inhibit its polymerization. Until

The effect of neutralization on the polymerization


rate was determined at three temperatures55,
70, and 857C. Both neutralization and temperature had a marked effect on the polymerization
rate, as shown in Figures 68. The rate of polymerization at 557C decreased with the increasing
level of neutralization up to 100% neutralization,
above which the rate increased again. At 125%
neutralization, the rate of polymerization was
about the same as that at 0% neutralization.
There are several potential explanations for the
decreasing rate with neutralization. These include a disruption of the template effect discussed
earlier, a lower inherent reactivity of the sodium
acrylate relative to acrylic acid, or an increase
in the ionic repulsion of the monomer from the
growing polymer due to the increased fraction of
ionized monomer. The latter explanation seems
most likely.
One very interesting observation was that the
more highly neutralized (e.g., 125%) feed batches
had virtually no inhibition times. Polymerization
started almost immediately, frequently at room
temperature during deoxygenation. However, the
reactions at 125% neutralization reached a plateau at lower conversion at elevated tempera-

Figure 4. Effect of persulfate concentration on the polymerization rate of 65% neutralized acrylic acid at 557C.

/ 8Q31$$5007

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

2034

ET AL.
CUTIE

Figure 5. The dependence of the initial polymerization rate on the square root of the
sodium persulfate concentration.

tures, possibly due to restricted mobility at this


high degree of neutralization.
Figure 9 shows the NMR spectra of polymers
at about 50% conversion, polymerized at different
levels of neutralization. The lines broaden dramatically at higher % neutralization, the source
of the broadening being due to restricted mobility
of the polymer. These findings support the idea
that restricted motion of the polymer backbone
at the higher levels of neutralization cause the
reactions to reach a plateau at lower conversions.
The initial rate of polymerization showed similar, but not identical, pH dependences at all temperatures. The rate of polymerization at 557C
showed a minimum at 100% neutralization,
whereas at 70 and 857C the minima were at 50

and 75% neutralization, respectively. The rates of


polymerization are given in Table I and Figure
10. For these experiments, initial monomer and
initiator concentrations were kept constant;
therefore, we have simply used the initial polymerization rates for our analyses. Figure 11 gives
an activation plot for these rate data. The activation energies, from the initial slopes of the plots
in Figure 11, are given in Table II. These plots
are fairly linear for all the levels of neutralization.
The activation energies for the initial rates of polymerization increased as the level of neutralization increased up to 100% neutralization. At 125%
neutralization, the activation energy decreased
again. These trends in activation energy suggest a
transition in mechanism at 100% neutralization.

Figure 6. The effect of % neutralization on the rate of polymerization of acrylic acid


at 557C.

/ 8Q31$$5007

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

2035

Figure 7. The effect of % neutralization on the rate of polymerization of acrylic acid


at 707C.

The increase in activation energy can be rationalized using the same arguments as for the pH dependence of the polymerization rates. As neutralization increases, the ionic repulsion between the
growing polymer in the ionized state and the ionized monomers also increases. This repulsion
would be maximized at 100% neutralization and
then be mitigated as the ionic strength of the medium increased due to continued neutralization.
The NMR kinetic data support an expression
for the rate of polymerization, Rp , as a product of
[M] 3 / 2 and [I] 1 / 2 . Table III shows values of Rp /

[M] 3 / 2 r [I] 1 / 2 determined in a similar range of


conditions from references previously cited along
with values determined in this work. No previous
studies were run at the 4M concentrations used
commercially. The dramatic effect of pH on the
rate, ranging from a factor of 50 (Kabanov) to
20 (Ito) for maximum/minimum, appeared to be
greater in the more dilute systems; our work
showed a range of about 5 at 557C (Table III).
Overall, activation energy values from previous
literature for the polymerization of sodium acrylate were 16.7 kcal /mol reported by Ito et al.

Figure 8. The effect of % neutralization on the rate of polymerization of acrylic acid


at 857C.

/ 8Q31$$5007

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

2036

ET AL.
CUTIE

Molecular Weight of Polyacrylic Acid

Temperature Effects

Figure 9. The effect of % neutralization on the 1HNMR line width of the Resonances of acrylic acid at
557C.

and 23.5 kcal /mol by Manickam. These may be


compared to the values of this work, given in
Table II.

Counterion Dependence
The dependence of the polymerization rate on
counterion was determined by comparing the rate
for acrylic acid neutralized with Na2CO3 to that
neutralized with K2CO3 . The samples were 75%
neutralized and the polymerization was run at
557C. These data, shown in Figure 12, indicate no
detectable difference in rate for these counterions.
These results are quite different from those discussed earlier for acrylamide polymerization.12
Even though no counterion dependence was observed for the polymerization rate, ionic strength
is known to have a marked effect on the rate.4,5
The addition of sodium chloride caused a notable
enhancement of the polymerization rate with a
concomitant increase in the molecular weight of
the polymer.

/ 8Q31$$5007

07-28-97 20:31:22

The effect of temperature on the molecular weight


of 65% neutralized polyacrylic acid was determined
at three temperatures (55, 70, and 857C). A mixture was prepared at 65% neutralization, as before,
in a glass vessel from sodium carbonate, acrylic
acid, sodium persulfate (1600 ppm based on acrylic
acid), and water to a solids content of 32% (4.1M).
Small microampoules (1.8 mm o.d.), described previously, were loaded with the feed mixture, deoxygenated for 45 min, sealed to exclude oxygen, and
placed in a water bath. At 557C, samples were withdrawn over a 90-min period, at 707C samples were
taken over 20- and 60-min periods, and at 857C
samples were taken over a 10-min period. The conversion at each time and temperature was determined in parallel experiments by polymerizing the
feed mixture while monitoring conversion by proton
NMR. The samples were hydrolyzed in dilute caustic and the molecular weight determined by SEC
(Fig. 13). Very high molecular weight polymer (5
million Daltons) was formed early in the polymerization, while the number average MW decreased
with conversion (Table IV, Fig. 14). Because of the
lack of molecular weight standards above 1,400,000
Daltons, the absolute values for polymers of greater
molecular weight should be used with caution and
only for relative comparison within a given study.

Effect of Initiator Concentration


on Molecular Weight
A 65% neutralized acrylic acid mixture was
spiked with increasing amounts of sodium persulfate and samples polymerized at 557C while monitoring conversion by proton NMR. Each sample
was polymerized for a set period of time to obtain
kinetic data and then the sample was analyzed
for molecular weight by SEC. Concentrations of
Table I. The Effect of Temperature and %
Neutralization on the Initial Rates of Polymerization
of Acrylic Acid
% Neutralization
Temperature

0%

25%

50%

75%

100%

125%

557C
707C
857C

4.0
7.2
14.

2.8
3.8
10.

1.5
2.4
6.2

1.2
2.4
6.8

0.83
2.9
9.2

3.0
6.5
10.

The units for the rates are % conversion per minute.

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

2037

Figure 10. The effect of temperature and % neutralization on the initial rate of acrylic
acid polymerization, circles557C, diamonds707C, and squares857C.

sodium persulfate ranging from 2.73 1 10 04 M


(218 ppm based on acrylic acid, 32% solids) to
8.72 1 10 03 M (6963 ppm based on acrylic acid)
were used (Table V). The conversion of the samples was not constant for each sample, which will
affect the final MW; however, a plot of the number
average molecular weight vs. 1/[sodium persulfate] 0.5 results in a reasonably good straight line
relationship (Fig. 14) similar to the rate of polymerization data presented in Figure 5. (This plot
does not go through the origin for the same reason
as that described for the plot of Figure 5.) An overlay of the SEC traces for the polymers presented
in Table VI is shown in Figure 15. The significant
influence of the initiator content on the polymer
molecular weight can be seen from these data. Of
course, the molecular weight of the polymer at

low conversion is normally used for determining


this relationship.

Effect of Neutralization on the Molecular


Weight of Polyacrylic Acid
The final polymer samples that resulted from the
NMR kinetic study of percent neutralization (pH)
effects on polymerization rate were hydrolyzed
with dilute caustic and analyzed by SEC to determine molecular weight.15 Samples prepared at
three temperatures (55, 70, and 857C) were characterized (Table VI). The neutralization of the
acrylic acid feed mixtures varied from 0 to 125%.
In contrast to the data at 65% neutralization presented in Table IV, differences in the final MW at
the three temperatures were observed at each
level of neutralization. The conflicting results may
have been the result of residual head space oxygen
in the samples prepared in the water bath polymerizations at 557C (Table IV), which could have
Table II. The Effect of Temperature and %
Neutralization on the Activation Energy for
Acrylic Acid Polymerization

Figure 11. Arrhenius plot for the polymerization of


acrylic acid at various degrees of neutralization.

/ 8Q31$$5007

07-28-97 20:31:22

polpa

% Neutralization

Ea (Kcal/mol)

0
25
50
75
100
125

9.6
9.8
10.6
13.6
18.4
9.2

W: Poly Physics

9605007

2038

ET AL.
CUTIE

Table III. Comparison of Polymerization Rates with Literature Values

Initiatora

liter/mol min
Rp
[M]3/2 [I]1/2

Reference

Temp.

pH

[M]

Manickam et al.5
Manickam et al.5
Manickam et al.5

50
50
50

1
4.2
11

0.138
0.172
0.258

KPS
KPS
KPS

0.01
0.01
0.01

0.4910.586
0.236
0.0910.444b

Kabanov et al.2,3
Kabanov et al.2,3
Kabanov et al.2,3

60
60
60

1.2
1.2
1.2

AIBN
AIBN
AIBN

0.005
0.005
0.005

0.14
0.19
0.00278

Ito et al.1
Ito et al.1
Ito et al.1

50
50
50

2.4
4.7
7.2

0.5
0.5
0.5

APS
APS
APS

0.00285
0.00285
0.00285

0.90
0.14
0.045

This work

55

4.5

NPS

0.0023

a
b

1
10
7

1.144.56

[I]

0.1260.239

KPS: potassium persulfate; NPS: sodium persulfate; APS: ammonium persulfate; AIBN: azo bis-isobutyronitrile.
Increasing ionic strength.

reduced the molecular weight of the samples or


simply be the result of the greater influence of
temperature on molecular weight at the lower %
solids (2.8M in this study vs. 4.1M in the study
reported in Table IV). The data in Table V were
determined at 65% neutralization, where the molecular weight differences at the three temperatures in (Table VI) are minimized. The overall
MW of the polymer samples in Table VI, prepared
at the lower % solids, are significantly less than
those in Table IV prepared at the higher % solids.
The highest molecular weight at all temperatures
is prepared at 0% neutralization, while the lowest
MW occurs when polymerizing at 100% neutralization. The molecular weight increases again
when polymerizing in the presence of excess sodium carbonate (125% neutralization). The MW
results shown graphically in Figure 16 follow the

Figure 12. The dependence of the polymerization


rate for 75% neutralized acrylic acid on counterion, sodium (triangles) vs. potassium (filled circles) at 557C.

/ 8Q31$$5007

07-28-97 20:31:22

same trend as the rate of polymerization vs. %


neutralization (Fig. 10).

Effect of Monomer Concentration


on Molecular Weight
Two experiments were conducted to determine
the effect of monomer concentration on the molecular weight of 65% neutralized acrylic acid. The
first study involved measuring the molecular
weight of the polymers by SEC that were prepared
during the NMR rate studies. The molecular
weight data are shown in Table VII. The molecu-

Figure 13. Number-average molecular weight of polyacrylic acid prepared at 65% neutralization and
31.9% solids (4.1 M) vs. conversion at 557C (squares),
707C (filled triangles), and 857C (circles).

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

2039

Table IV. Molecular Weight of 65% Neutralized Polyacrylic Acid as a Function of Polymerization Time and
Temperature
Temp (7C)

Time (minutes)

Mw 106

Mn 106

Mp 106

% Conv.a

55
55
55
55
55
55
55
55
55
55
70
70
70
70
70
70
70
70
70
70
70
70
85
85
85
85
85
85

5
8
10
12
16
20
30
40
60
90
4
6
8
10
20
4
6
8
10
10
40
60
1
2
3
4
6
10

8.135
6.604
6.898
5.964
6.963
4.215
3.276
3.527
1.930
1.316
5.896
4.393
4.257
4.934
4.591
10.690
11.104
9.858
1.1392
9.711
6.269
6.373
6.142
6.238
4.612
4.817
4.105
3.208

2.100
2.033
1.639
1.636
1.819
1.058
0872
0.917
0.480
0.359
1.912
1.210
1.355
1.109
1.327
1.536
1.354
1.713
1.510
1.349
0.708
0.627
1.779
2.091
1.219
1.245
0.863
0.555

10.967
6.147
6.880
5.264
6.382
3.418
2.290
2.290
1.314
0.845
5.571
2.310
2.495
3.143
2.355
7.557
8.730
5.540
7.557
5.696
2.195
2.833
5.786
6.123
2.855
2.967
2.447
1.663

2
7
11
14
22
25
43
62
80
92
8
11
19
26
59
8
11
19
26
59
80
90
3
15
30
44
64
81

Estimated from time of polymerization using separate NMR experiments.

lar weight drops significantly with decreasing


monomer concentration, but with significant scatter. Ampoules containing the same feed mixtures
were also polymerized in a water bath for various
times and the samples analyzed for molecular
weight by SEC. These data are presented in Table
VIII and plotted in Figure 17. Based on the slope
of the data in Figure 17, the number-average molecular weight is proportional to approximately
the 1.5 power of the monomer concentration, consistent with theoretical prediction.

Kinetic Model

Figure 14. The effect of sodium persulfate concentration on the number-average molecular weight of 65%
neutralized polyacrylic acid prepared at 557C.

/ 8Q31$$5007

07-28-97 20:31:22

The simplest polymerization model, which has a


first-order dependence in monomer and half order
in initiator, does not adequately fit the data presented here, as noted earlier. Using the arguments developed so far the following kinetic
scheme is proposed:

polpa

W: Poly Physics

9605007

2040

ET AL.
CUTIE

Table V. Molecular Weight of 65% Neutralized Polyacrylic Acidd Samples Prepared at 31.9% Solids in NMR
Tubes at 557C
[Na2S2O8] mol/liter

Mwc 106

Mnc 106

Mpc 106

Final % Conversion

0.000273a
0.0016a
0.00433b
0.00872b

12.019
3.897
0.208
0.087

1.824
0.605
0.065
0.033

17.126
2.129
0.164
0.0613

44
78
82
92

Used 5 mm NMR tube.


Used 2 mm NMR tube.
Molecular weight affected by the conversion of the samples.
d
Acrylic acid contained 177 ppm MEHQ.
b
c

I r (2R c )

kd

(2)

(2R c ) / M r RM / R k*

(3)

R / M r RM
RM / M r RM

RM / RM r polymer

k9

(4)

kp

(5)

kt .

(6)

In this scheme (2R c ) represents two radical


species trapped in a solvent cage. Only when these
caged radicals are intercepted by a monomer does
polymerization ensue. This is essentially the concept proposed by Ishige and Hamieliec 14 for acrylamide.
The reaction scheme set forth above results in
the following kinetic expressions:

d[M]
kp[RM ][M].
dt

(7)

The concentration of the caged radicals [(2R c )]


d[(2R c )]
kd[I] 0 k*[(2R c )][M].
dt

(8)

Now if these rates were comparable (the caged


radicals are in steady state) and this differential
were equal to zero, then the expression for the
overall rate of polymerization would collapse into
its normal form, and there would be a first-power
dependence on monomer. In other words, if the

Table VI. Molecular Weight of Polymers Made at Varied % Neutralization and Temperatures
Temp (7C)

% Neut

% Conver.

Mw 106

Mn 106

Mp 106

55
55
55
55
55
55
70
70
70
70
70
70
85
85
85
85
85
85

0
25
50
75
100
125
0
25
50
75
100
125
0
25
50
75
100
125

84
82
84
78
62
85
89
86
79
77
85
65
82
90
83
84
74
90

3.449
2.048
1.637
1.402
1.947
1.600
2.084
1.113
0.975
1.469
0.871
1.162
1.270
0.805
0.869
1.490
0.643
1.340

0.563
0.458
0.328
0.303
0.278
0.352
0.264
0.227
0.228
0.205
0.151
0.264
0.207
0.154
0.168
0.163
0.101
0.201

1.469
1.252
1.026
0.858
0.986
0.967
1.179
0.693
0.555
0.640
0.380
0.693
0.621
0.465
0.363
0.537
0.291
0.505

/ 8Q31$$5007

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

2041

Figure 15. Molecular weight distribution curves of polyacrylic acid samples prepared
at 557C with varied sodium persulfate levels as listed in Table V.

rate controlling step is the decomposition of initiator instead of the interception of the caged radicals by monomer, then no unusual dependence
on monomer concentration would be discernible.
Because this explanation is not consistent with
the data (Fig. 2), it is concluded that kd[I] must
be considerably larger than k* [((2R c )][M].
One approach is to assume there is a relatively
constant fraction of caged radicals proportional to
the decomposed initiator. Manickam5 proposed a
direct reaction between monomer and persulfate.
Using the relation
[(2R c )] fkd[I]

d[RM ]
2k*[M] fkd[I] 0 2kt[RM ] 2 . (10)
dt
Also assuming steady state for [RM ] and [R ],
then the growing radical concentration is

[RM ]

Rp kp[M][RM ]
kp[k*f kd /kt ]

where f is an efficiency factor encompassing a constant fraction of caged radicals, then

/ 8Q31$$5007

07-28-97 20:31:22

(11)

which is proportional to [I] 1 / 2 and [M] 1 / 2 and


leads to a propagation expression

(9)

Figure 16. Effect of temperature and neutralization


on the molecular weight of polyacrylic acid, 557C
(squares), 707C (diamonds), and 857C (circles) (a,b).

k*f kd[I][M]
kt

(12)

1/2

[M]

3/2

[I]

1/2

(13)

This can explain the concentration dependence

Figure 17. Effect of acrylic acid concentration on the


number-average molecular weight of 65% neutralized
polyacrylic acid polymerized at 557C and containing
1600 PPM sodium persulfate (B.O.A.A.)

polpa

W: Poly Physics

9605007

2042

ET AL.
CUTIE

Table VII. Effect of Monomer Concentration on the Molecular Weighta of 65% Neutralized Polyacrylic Acid
[Monomer] mol/liter

% Conv.

Mw 106

Mn 106

Mp 106

Tube Diameter (mm)

4.56
2.28
1.71
1.14

65
60
70
80

6.802
2.576
0.943
1.170

1.193
0.549
0.226
0.274

4.212
1.755
0.709
0.741

2
5
5
5

Analysis of samples polymerized in the NMR.

of the rate curves in a general way. The effect of


pH, ionic strength, and temperature can be incorporated as empirical values for an overall coefficient. It is also important to take into account
the variation in kd , the decomposition constant
for sodium persulfate, as a function of these same
parameters. The temperature dependence for the
kd has been determined under these conditions,
and this dependence is reflected in the plots in
Figure 11.

CONCLUSIONS
The polymerization rate, as well as the molecular weight, showed a dependence on the 32 power
of monomer concentration for [ M ]0 values up
to 4.5 molar. This dependence was accounted
for by interaction of the monomer with either
the initiator itself or caged radicals resulting
from it. In either case, the steady-state concentration of radicals appeared to depend on the
square root of monomer concentration, leading
to the overall [ M ] 3 / 2 dependence for R p . Both
R p and molecular weight possessed a 12 power
dependence on the initiator ( sodium persulfate ) concentration. Values for Rp / [ M ] 3 / 2[ I ] 1 /
2
obtained from the literature for acrylic acid
polymerization, all of which were from experiments at somewhat lower monomer concentra-

tions, varied by a factor of 10. Values obtained


in this work at higher monomer concentrations
were within this range. A minimum rate of polymerization was observed near neutral pH, as
was recognized previously. This minimum appeared be somewhat temperature dependent,
resulting in a variation in overall activation energies with pH. In addition, the rate differences
observed for unneutralized polymerizations vs.
the minimum rates were smaller than those reported in the literature. This was perhaps due
to the higher monomer concentrations of this
study.
The initial rate of polymerization of 125% neutralized samples was very high, but reached a plateau at 5060% conversion. This appeared to be
due to a mobility problem at the higher monomer
(polyelectrolyte) concentrations in this study. The
activation energies for acrylic acid polymerization
between 0 and 125% neutralization were found to
range from 9.2 to 18.4 kcal/mol, with a maximum
at 100% neutralization. This large change in activation energy at 100% neutralization probably resulted from a change in the polymerization mechanism.
No significant effect of K / vs. Na / counterions
was observed on the rate of polymerization of neutralized polyacrylic acid. Molecular weights varied with percent neutralization showing a minimum at 100% neutralization and higher molecular weights at 0 and 125% neutralization.

Table VIII. Effect of Monomer Concentration on the Molecular Weighta of 65% Neutralized Polyacrylic Acid
[Monomer] mol/liter

Approx. % Conv.

Reaction Time (min)

Mw 106

Mn 106

Mp 106

4.56
3.80
2.28
1.14

65
60
70
63

45
55
130
390

8.550
5.896
5.102
0.916

1.107
0.734
0.563
0.122

3.783
2.308
2.036
0.724

Analysis of samples polymerized in a water bath at 557C.

/ 8Q31$$5007

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

APPENDIX
Raw Data for the Rates of Polymerization
Raw Data for Figure 2, the Rate of Polymerization of 65% Neutralized Acrylic Acid at
557C as a Function of Acrylic Acid Concentration (Fractional Conversion vs. Time)
Time (min)

4.56 Molar

2
4
6
8
10
12
14
16
18
20
24
28
32
36
40
44
48
56
64
72
80
88
96
104
112
120
128
136
144
152
160
168
192
224
256
320

0.05
0.07
0.13
0.17
0.21
0.25
0.30
0.34
0.39
0.45
0.53
0.61
0.66
0.72
0.77
0.81

2.29 Molar
0

1.17 Molar

0.02

0
0
0
0

0.14

0.05

0.24

0.14

0.32

0.21

0.42

0.28

0.46
0.51
0.56
0.6
0.64
0.66
0.7
0.72
0.74
0.75
0.76
0.77
0.79
0.79
0.8
0.81
0.83

0.32
0.37
0.4
0.43
0.46
0.49
0.5
0.55
0.58
0.61
0.64
0.65
0.66
0.7
0.72
0.75

The Initial Monomer Concentration and Initial Rate Data for the Samples of Figure 3
Rp (%/min)

[Monomer]

Rp (m/l/min)

log[monomer]

log Rp (m/l/min)

2.2
2
1.2
1.9
0.65
1.12
1.83
1.91
1.26

4.56
3.08
2.3
3.8
1.17
2.29
3.41
4.56
2.28

0.1
0.0616
0.0276
0.0722
0.007605
0.0256
0.0624
0.0871
0.0287

0.659
0.4886
0.3617
0.5798
0.0682
0.3598
0.5328
0.659
0.3579

01
01.21
01.559
01.141
02.119
01.592
01.205
01.06
01.542

/ 8Q31$$5007

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

2043

2044

ET AL.
CUTIE

APPENDIX Continued
The Initial Monomer Concentration and Initial Rate Data for the Samples of Figure 3
Rp (%/min)

[Monomer]

Rp (m/l/min)

log[monomer]

log Rp (m/l/min)

1.22
0.92

1.71
1.14

0.02086
0.0105

0.233
0.0569

01.681
01.979

Raw Data for Figure 4, Effect of Persulfate Concentration on the Polymerization


Rate of 65% Neutralized Acrylic Acid at 557C (Fractional Conversion vs. Time)
Time (min)
2
3
4
5
6
7
8
9
10
12
14
16
18
20
24
28
32
36
40
44
48
52
56
60
64
72
80
88
96
112
128
144

0.273 mmolar

1.60 mmolar

4.33 mmolar

8.72 mmolar

0
0
0
0
0.02

0
0
0.03
0.04
0.07
0.09
0.15
0.18
0.23
0.31
0.39
0.46
0.52
0.57
0.65
0.74
0.76

0
0.04
0.06
0.09
0.16
0.21
0.28
0.36
0.41
0.49
0.61
0.69
0.72
0.76
0.79
0.85
0.87
0.9
0.92

0
0
0
0
0
0
0
0
0
0
0
0
0
0.04
0.04
0.05
0.05
0.07

0.08
0.12
0.16
0.22
0.27
0.32
0.4
0.49
0.55
0.62
0.65

0.09

0.71

0.13

0.78

0.03

0.82

0.14
0.15
0.19
0.23
0.24
0.31
0.38
0.44

Raw Data for Figure 6, the Effect of % Neutralization on the Rate of Polymerization
of Acrylic Acid at 557C (Fractional Conversion vs. Time)
Time (min)

0%

25%

2
4
6
8
10
12

0
0.03
0.1
0.18
0.26
0.32

/ 8Q31$$5007

50%

75%

100%

0
0
0
0.03
0.1

07-28-97 20:31:22

0.06

0.05

polpa

W: Poly Physics

125%
0.07
0.11
0.13
0.22
0.28
0.34

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

APPENDIX Continued
Raw Data for Figure 6, the Effect of % Neutralization on the Rate of Polymerization
of Acrylic Acid at 557C (Fractional Conversion vs. Time)
Time (min)

0%

25%

14
16
18
20
22
24
28
32
36
40
44
48
50
52
56
60
64
68
72
80
84
96
112
120
124
128
144
160
176
192
200

0.42
0.47
0.52
0.57
0.61
0.68
0.74
0.76
0.79
0.81
0.84
0.84

0.13
0.22
0.25
0.3
0.36
0.43
0.55
0.61
0.66
0.7
0.73
0.75
0.78
0.79
0.82

50%

75%

100%

0.13

0.09

0.18

0.14

0.24
0.31
0.37
0.41
0.47
0.5
0.56
0.62

0.19
0.24
0.29
0.35
0.38
0.44
0.47

0.62

0.52

0.68

0.58

0.16

0.86

0.71
0.74

0.62
0.66

0.24

0.85
0.85
0.86

0.8
0.83
0.84

0.73
0.76
0.78
0.78

0.07
0.08

0.1

125%
0.41
0.46
0.51
0.56
0.58
0.62
0.67
0.72
0.74
0.76
0.78
0.8

0.83

0.3
0.37

0.43
0.54
0.62
0.7
0.78
0.79

Raw Data for Figure 7, the Effect of % Neutralization on the Rate of Polymerization
of Acrylic Acid at 707C (Fractional Conversion vs. Time)
Time (min)

0%

25%

50%

75%

100%

125%

2
4
6
8
10
12
14
16
20
24
28
32
36
40
44

0.12
0.35
0.41
0.54
0.64
0.7

0
0.06
0.14
0.2
0.29
0.37
0.46
0.51
0.61
0.68
0.73
0.77
0.81
0.83
0.86

0
0.04
0.09
0.15
0.21
0.22

0
0.05
0.09
0.13
0.17
0.22
0.28
0.3
0.39
0.47
0.51
0.57
0.6
0.65
0.67

0
0
0.08
0.14

0.13
0.32
0.39
0.52

0.24

0.57
0.57
0.59
0.61
0.65
0.64
0.64

/ 8Q31$$5007

0.79
0.84
0.87
0.89

07-28-97 20:31:22

0.32
0.43
0.47
0.54
0.57
0.61
0.65
0.68

polpa

0.37
0.5
0.6
0.67
0.73
0.75
0.81
0.81

W: Poly Physics

0.64

9605007

2045

2046

ET AL.
CUTIE

APPENDIX Continued
Raw Data for Figure 7, the Effect of % Neutralization on the Rate of Polymerization
of Acrylic Acid at 707C (Fractional Conversion vs. Time)
Time (min)

0%

25%

48
52
56
60
64
68
72
80

50%

75%

100%

125%

0.71
0.73
0.77

0.71
0.71
0.74
0.77

0.84
0.88
0.85

0.65

0.77
0.79

Raw Data for Figure 8, the Effect of % Neutralization on the Rate of Polymerization
of Acrylic Acid at 857C (Fractional Conversion vs. Time)
Time (min)
1
2
3
4
5
6
7
8
9
10
12
14
16
18
20
22
24
26
28
32
36
40

0%

25%

50%

75%

0.05

0.04

0.31

0.15

0.05

0.04

0.54

0.35

0.18

0.18

0.68

0.52

0.31

0.34

0.75
0.85
0.82

0.67
0.74
0.78
0.81
0.83
0.87
0.88
0.9

0.43
0.53
0.62
0.65
0.69

0.43
0.52
0.58
0.64
0.68
0.7
0.75
0.79
0.82
0.82
0.84

0.77
0.79
0.79
0.83

100%

125%

0
0.07
0.16
0.28
0.31
0.49
0.56
0.62
0.63
0.65
0.68
0.72
0.7
0.74
0.76
0.72
0.73
0.74
0.72
0.74
0.74
0.74

0
0
0.07
0.2
0.36
0.52
0.61
0.71
0.74
0.8
0.84
0.86
0.87
0.88
0.88
0.9

Raw Data for Figure 12, the Dependence of the Polymerization Rate for 75%
Neutralized Acrylic Acid on Counterion, Sodium vs. Potassium at 557C (Fractional
Conversion vs. Time)
Time (min)
4
6
8
10
12
16
20
24
28

/ 8Q31$$5007

Potassium

Sodium

0
0.01
0.08
0.12
0.18
0.26
0.35
0.42
0.46

0.04
0.12
0.23
0.31
0.38
0.45

07-28-97 20:31:22

polpa

W: Poly Physics

9605007

ACRYLIC ACID POLYMERIZATION KINETICS

2047

APPENDIX Continued
Raw Data for Figure 12, the Dependence of the Polymerization Rate for 75%
Neutralized Acrylic Acid on Counterion, Sodium vs. Potassium at 557C (Fractional
Conversion vs. Time)
Time (min)

Potassium

Sodium

32
36
40
48
56
60
64
72
80
96
104
112
120
128
144
160

0.48
0.55
0.56
0.63
0.67

0.54

0.7

0.74
0.75
0.79

0.78
0.8
0.83
0.86
0.87
0.88

REFERENCES AND NOTES


1. H. Ito, S. Shimizu, and S. Suzuki, J. Chem. Soc.,
Jpn., Ind. Chem. Sect., 58, 194 (1955).
2. V. A. Kabanov, D. A. Topchiev, T. M. Karaputadze,
and L. A. Mkrtchian, Eur. Polym. J., 11, 153
(1975).
3. V. A. Kabanov, D. A. Topchiev, and T. M. Karaputadze, J. Polym. Sci., 42, 173 (1973).
4. S. P. Manickam, K. Venkatarao, U. Chandra Singh,
and N. R. Subbaratnam, J. Polym. Sci., Polym.
Chem. Ed., 16, 2701 (1978).
5. S. P. Manickam, K. Venkatarao, and N. R. Subbaratnam, Eur. Polym. J., 16, 483 (1979).
6. F. Laborie, J Polym. Sci., Polym. Chem. Ed., 15,
1255 (1977).
7. F. Laborie, J Polym. Sci., Polym. Chem. Ed., 15,
1275 (1977).

/ 8Q31$$5007

0.6
0.65

07-28-97 20:31:22

8. A. Chapiro and J. Dulieu, Eur. Polym. J., 13, 563


(1977).
9. N. I. Galperina, T. A. Gugunava, V. F. Gromov,
P. M. Khomikovskii, and A. D. Abkin, A. D. Vysokomol. Soyed., 7, 1455 (1975).
10. J. P. Riggs and F. Rodriguez, J. Polym. Sci., Part
A-1, 5, 3167 (1967).
11. S. Lenka, P. L. Nayak, S. B. Dash, and S. Ray, Coll.
Polym. Sci., 261, 40 (1983).
12. R. Praba and U. S. Nandi, J. Polym. Sci., Polym.
Chem. Ed., 15, 1973 (1977).
13. S. S. Cutie, D. E. Henton, C. Powell, R. E. Reim,
P. B. Smith, and T. L. Staples, J. Appl. Polym. Sci.,
64, 577 (1997).
14. T. Ishige and A. E. Hamielec, J. Appl. Polym. Sci.,
17, 1479 (1973).
15. S. S. Cutie and S. J. Martin, J. Appl. Polym. Sci.,
55, 605 (1995).

polpa

W: Poly Physics

9605007

Você também pode gostar