You are on page 1of 8

Dissolution of polypropylene in organic

solvents: 1. Partial dissolution


D. A. Blackadder and G. J. Le Poidevin*
Department of Chemical Engineering, University of Cambridge, Cambridge CB2 3RA, UK
(Received 8 January 1976)

The dissolution of polypropylene in organic solvents has been investigated in detail. The complete
study will provide an interesting comparison with the literature concerning the dissolution of amor-
phous polymers. Under appropriate conditions polypropylene dissolves at a steady rate, but there is
a preliminary induction period. The results reported here concern this period, and provide the foun-
dation for later parts of the work. It was necessary to study solvent sorption by polymer as well as
the partial or fractional dissolution characteristic of the induction period. The Hildebrand solubility
parameter proved useful in understanding the behaviour of polypropylene soaked in various solvents
over a range of temperatures. Factors controlling the partial dissolution have been identified and
analysed.

INTRODUCTION tots are of importance, as for polypropylene, the extracted


In a series of papers 1-7 Ueberreiter and his collaborators material may consist mainly of low molecular weight iso-
have provided a detailed description of the dissolution of tactic chains and also of atactic chains, not necessarily of
polystyrene, a typical amorphous polymer, in organic sol- low molecular weight 9-11.
vents. Among the factors identified as having a bearing on At temperatures well below their melting points, semi-
the rate of dissolution were: the nature and molecular crystalline polymers do not dissolve wholly in organic sol-
weight of the polymer; the viscosity and size of the solvent vents. This is because the positive entropy changes asso-
molecules; the comparability of polymer and solvent; and ciated with fusion and mixing do not sufficiently offset the
the temperature and conditions of agitation in the system. large heat of fusion and the overall free energy change for
Hitherto, relatively little has been reported concerning the crystallite dissolution is therefore positive. However, even
dissolution of semi-crystalline polymers where one would at low temperatures, most solvents can penetrate the non-
expect the crystallinity to have some effect and, less ob- crystalline regions to some degree, resulting in the solva-
viously perhaps, where some material and some regions are tion of individual segments of polymer chains and overall
more soluble than the remainder under given conditions. swelling of the specimen. At the equilibrium swollen state,
The dissolution of polyprcpylene has now been studied in the osmotic pressure associated with the mixing of solvent
detail. and polymer is just balanced by the elastic retractive forces
The present paper, Part 1 of a series, describes the effect supplied by a network of interlamellar tie-molecules.
of soaking polypropylene film in organic solvents at various
temperatures up to 110°C. It will be shown in a later paper
that bulk specimens of this polymer can be made to dis-
solve at a constant rate at temperatures from 110°C up- EXPERIMENTAL
wards, once an induction period has been completed. This
induction period arises because the penetration of polymer Materials
by solvent is a prerequisite for dissolution and it takes time
to establish a situation in which characteristic events are Polymer. The polypropylene used was a homopolymer,
occurring at different depths of penetration. Clearly there GXM 43, kindly supplied by ICI Plastics Division. The
amount of atactic material present was estimated to be
will be a gradation in composition and properties between
dry bulk polymer to which no solvent has penetrated and about 2 wt %, and the polymer contained the usual small
the boundary layer of liquid adjacent to the swollen surface amounts of antioxidant and stabilizer. In view of the rela-
tively high temperatures used in some of the present experi-
of the solid.
In order to understand the induction period it is neces- ments these additives were welcome and no attempt was
made to remove them. The polymer was quoted as having
sary to investigate the general effects of solvents on the
a melt flow index of 2.16 and a melt viscosity of 5.9 × 103
polymer under conditions where total dissolution does not
occur. This involves measurement of the amount of poly- Nsec/m at 190°C.
mer dissolved and the amount of solvent imbibed under Solvents. The p-xylene was a 99% product obtained
various conditions, and the interpretation of the results in from ICI. The other organic solvents, of SR or AR grade,
the light of thermodynamics and kinetics. It is already were used as received.
known that fractional dissolution of polyethylene can occur
at temperatures well below the melting point 8 and the mate- Procedures
rial dissolved is of low molecular weight. When steric fac- Preparation of film specimens. To investigate the effect
of soaking polymer in solvent it is convenient to use film
* Now at the Electricity Council Research Centre, Capenhurst, rather than thicker specimens. Polypropylene chips were
Chester. UK. compression moulded between platens at 220°C-for 4 rain.

POLYMER, 1976, Vol 17, May 387


Dissolution of polypropylene in organic solvents (1): D. A. Blackadder and G. J. Le Poidevin
3.C completed the drying process by removing residual solvent.
The final mass of the specimen was used to calculate the
percentage of the original material which had been dissolved
out.
Density and crystallinity. Densities were measured in a
density gradient column at 25°C using p-xylene and chloro-
benzene as column liquids. Fragments of polymer t'dm
2.C
were cut so as to have a specific surface of less than 10 m2/
kg, as recommended by Blackadder and Keniry 13. In gene-
ral they reached equilibrium in the column in less than 24 h.
It can be shown that loss of soluble material to the column
liquids requires a small correction to the observed density,
which would otherwise be overestimated. With the correc-
I.O tion, values are reliable to 0.2 kg/m 3. When required, cry-
stallinities were calculated from densities using Natta's
formula14:

983 + 0.90(T - 93) - (106/d)


(1 - x ) = (1)
0.48(T - 93)
I I I I I I
16 18 20 where (1 - •) is the weight fraction of crystalline mate-
/~ (M j/m3) v2
rial, commonly written as Wc, and d is the density (kg/m 3)
Figure 1 Relationship between solubility parameter and equili- at T(K).
brium solvent uptake by polypropylene at 25oc: 13, chlorinated
solvents; O, aromatic and napthenic solvents; A, n-alkanes
RESULTS AND DISCUSSION
The resulting films were quench cooled by immersion in
cold water. The thickness was about 0.3 mm, and rectan-
Solubility of polypropylene in various solvents
gular pieces measuring 50 mm by 16 mm were cut as re-
quired. The average thickness of each strip was determined In order to interpret polymer solubility in thermodyna-
from ten micrometer measurements spaced along both of mic terms, it is natural to turn to well-tried parameters
the long sides. The density of strips was adjusted when which have enjoyed some success in characterizing simpler
necessary by annealing for 2 h in a vacuum oven at an systems. The strength of the non-valence or secondary
appropriate temperature. bonding in a liquid system is related to the cohesive energy
Solvent sorption. Several dry strips of fdm were weighed density, AEv/V, where AE v is the molar energy of vaporiza-
to 0.05 mg and then transferred simultaneously to tubes tion and V is the molar volume. It is better to predict and
of solvent held in a thermostat controlled to -+0.05°C. interpret solubilities in terms of the square root of the co-
After a suitable immersion time a strip was removed from hesive energy density, known as the Hildebrand solubility
the liquid, lightly pressed between tissues, and hung from parameter is given by:
one arm of a balance. A second stopwatch was started as
the strip was removed from the liquid and the total mass 8= (2)
of polymer and imbibed solvent was noted at 30 sec inter-
vals over a period of 3 min. A graph of mass against time
permitted extrapolation to the moment of removal, thus Liquids with like values of ~ tend to be equally good sol-
giving the true mass of solvent imbibed per unit mass of dry vents for a given solute and to be mutually compatible.
polymer after a known period of immersion 12. The time This leads to indirect methods of measuring 8 for a poly-
for attainment of equilibrium at 25°C varied from a few mer where AEv/V cannot be evaluated. Michaels et al. 16
hours to several days but all equilibrium values were based have measured the equilibrium sorption of various solvent
on the uptake after 7 days, using the original mass of dry vapours by polypropylene at 25°C, and found that the
polymer as a basis, irrespective of the amount of polymer equilibrium sorption (mass of solvent vapour sorbed by unit
dissolved while solvent was imbibed. At higher tempera- mass of dry polymer) reached a maximum for cyclohexane
tures it is quite possible that equilibrium was not always which had a solubility parameter of 16.8 (MJ/m3) 1/2. Since
attained, but the results were instructive nevertheless. The the thermodynamic dissolution temperature of the poly-
error in measuring the uptake was estimated to be only ~1% mer, Ts, was a minimum in the same solvent, it was sug-
at 25°C, rising to at least 5% at 100°C. It was often con- gested that polypropylene had a solubility parameter of
venient to convert the measured uptake into units of volume ~16.8 (M J/m3) 1/2. Other authors17 have pointed out that
sorbed per unit mass of polymer. the viscosity of solutions of a polymer in various solvents
Soluble fraction. Strips of fdm, as described above, at a given temperature and concentration shows a maxi-
were vacuum dried at 80vC for 12 h before weighing. For mum when the polymer and solvent have most nearly the
initial densities between 908 and 918 kg/m 3, this treatment same ~i value. The maximum arises because the polymer
was shown to cause no detectable change in density. Strips chains are most uncoiled in solvents of high comparability.
were soaked for 7 days in solvents at chosen temperatures, Figure 1 shows Q, the equilibrium volume of solvent
with a solvent: polymer ratio of 250:1 to avoid any pos- imbibed per unit mass of dry polypropylene at 25°C, plot-
sibility of saturating the solvent with soluble species. After ted against literature values 1s'18 of the solubility parameters
removal from solvent, specimens were wiped dry and left of the various solvents, calculated from the heats of vap-
in air for 24 h. A period of 12 h in a vacuum oven at 60°C orization and the molar volumes. The density of the poly-

388 POLYMER, 1976, Vol 17, May


Dissolution o f polypropylene in organic solvents (1): D. A. Blackadder and G. J. Le Poidevin
Table 1 Effect of solubility parameter, 6, on the equilibrium correlation could be established, though shorter soak times
swelling, Q, of polypropylene at 25°C might have shown up some dependence on crystallinity 12.
Q X 104 At least the results suggest that the material soluble at 25°C
Solvent Solvent type (m3/kg) &(M j/m3) 1/2 is essentially non-crystalline. The lowest recorded weight
loss was 0.11% in 1-chloronaphthalene and even this was
Benzene Aromatic 1.16 18.7 probably more than mere antioxidant removal. Further
Toluene Aromatic 1.24 18.2
p-Xylene Aromatic 1.30 17.9 experiments were devised to test the hypothesis that the
Cumene Aromatic 1.06 17.7 1.1% or so dissolved by p-xylene, the best solvent used,
Cyclohexane Hydroaromatic 2.20 16.8 was polymeric and mostly atactic.
Methylcyclohexane Hydroaromatic 1.80 16.1
Tetrahydro- Naphthenic 1.04 19.5
naphthalene
Decahydro- Naphthenic 1.91 17.2 4'0
naphthalene
a-Chloro- Chlorinated 0.74 21.1
naphthalene naphthenic
Chlorobenzene Chlorinated 1.22 19.9
hydrocarbon
Carbon Chlorinated 2.08 17.6
tetrachloride hydrocarbon
n-Dodecane n-Alkane 1.06 16.1 3.0
n-Decane n-Alkane 1.10 15.8
n-Octane n-Alkane 1.21 15.5

1
n-Heptane n-Alkane 1.66 15.1
n-Hexane n-Alkane 1.48 14.7

%. "%.
propylene used was 907.7 kg/m 3 at 25°C. In good agree- ~ 2.0
E
",C,-.,
ment with Michaels et aL 16 the maximum uptake of solvent
for the naphthenic and aromatic solvents occurs at a value "_o
of 5 between 16.5 and 16.8 (MJ/m3) 1/2. There is a sepa-
rate curve for n-alkanes, and signs of yet another for chlo-
rinated hydrocarbons. The swelling data appear in Table 1. \\
It is significant that for alkanes the maximum occurs at a \
1.0
lower 6 value than for the naphthenics and aromatics; Hilde-
brand 19 and other workers have noted anomalies in binary
solutions where one component is a paraffin. In particular
Hildebrand 2° evaluated 8 values by means of regular solu-
tion theory and showed that they were consistently higher
than values obtained from equation (2). He quoted iso-
I I I I I ~ I
octane and n-heptane as having 8 values 2.0 and 1.2 (MJ/ q4 16 18 2k,J
m3) 1/2 higher, respectively, than those determined from 8 (Ma/m3)V2
AE v and V, whereas the value for cyclohexane was much Figure 2 Relationship between solubility parameter, equilibrium
the same for both methods of evaluation. This reinforces solvent uptake and mass percentage loss of soluble material for
the argument that 8 for polypropylene is indeed very close polypropylene at 50°C: El, • chlorinated solvents; o, • aromatic and
to 16.8 (MJ/m3) 1/2, and there is a case for correcting alkane naphthenic solvents; A, • n-alkanes. Equilibrium solvent uptake
( ); mass percentage loss (-- -- --)
values by ~1.7 (MJ/m3) 1/2, an amount which would make
the maximum of the alkane curve on Figure 1 coincide
with that of the other curve. 1.2
It is noteworthy that cumene falls between the aroma- 0 (J
tic-naphthenic curve and the n-alkane curve. This may be
because cumene has a fairly bulky aliphatic side chain,
which confers some alkane character without being large
enough to overshadow the effect of the benzene ring. If
08
the abscissa of the point for cumene is indeed too small
it would account for the observed displacement.
Repeating the above experiments at 50°C gave the re-
suits shown on Figure 2, which includes data discussed in
the following section.
0-4
Soluble fraction
It was mentioned in the introduction that atactic and
some low molecular weight isotactic material might be ex-
pected to dissolve at temperatures substantially lower than
I / I I I I I
the dissolution temperature of the bulk polymer. An 60 I00 140 180
attempt was made to correlate the densities of 3 distinct tih)
polypropylene specimens with the weight losses recorded Figure 3 Loss of soluble material (wt%) plotted against time for
after soaking in 4 different solvents for 7 days at 25°C. No polypropylene film density 908 kg/m 3 immersed in p-xylene at 25°C

P O L Y M E R , 1976, V o l 17, May 389


Dissolution of polypropylene in organic solvents (1): D. A. Blackadder and G. J. Le Poidevin
30 the dissolved material had an in situ density of between 848
and 860 k~/m 3, limits which enclose Natta's value 14 of
857 kg/m~for the density of wholly amorphous polypropy-
lene at 25°C. Since the atactic content of the polymer is
~2%, it seems that the 1.12% lost to p-xylene in 7 days is

_20 o very probably wholly atactic.


Having demonstrated that at 25°C the overall density
of a polypropylene specimen increased as soluble material
was leached from the non-crystalline regions, it was appro-
priate to investigate higher temperatures and other solvents,
always with a view to understanding the induction period
preceding steady total dissolution at higher temperatures
still. Figure 2 shows that at 50°C there is a rather close
IO correspondence between solvent untake and weight loss
in the same period, and the solubility parameter correlates
jo g both properties effectively, bearing in mind the possible
corrections to the 6 values for some solvents. Figure 4
0
0 0 gives the measured weight loss for various temperatures
and solvents plotted against the appropriate density after
I 0 I I I
I
soaking for the usual 7 days. The intercept corresponds to
07 911 915 919 unsoaked film and the solid curve is the relationship ex-
ds(kglm 3) pected if: (a) all the material dissolved had the density of
Figure 4 Loss of soluble material (wt%) plotted against density amorphous polypropylene; and (b) the soaking treatment
of polypropylene after treatment with various solvents (initial den- left the remainder totally unaffected. The curve was ob-
sity 908 kg/m3): ©, measured values;, , calculated values assum- tained as follows. Equation (1) relates (1 - k)u to du as
ing soluble material is leached only from non-crystalline regions of well as (1 - k)s to ds, where the subscripts u and s denote
the bulk polymer unsoaked and soaked specimens. If the mass of crystalline
material is indeed unaffected by any soaking treatment
Figure 3 shows one consequence of soaking polypropy- then it follows that:
lene f'dm in p-xylene at 25°C. Several selected strips of
uniform thickness (0.300 -+ 0.010 mm) and density 907.7 (1 - X ) u x 100
= (1 - x ) s (3)
kg/m 3 were weighed and separately immersed in p-xylene. 100 - m
Samples were removed after appropriate times and dried
as described earlier, so that the soluble fraction might be Evidently the percentage loss in weight, m, can be calcu-
calculated at various intervals over a period of 7 days. After lated from this expression ifds and du can be measured.
a steep initial rise the curve flattened off to become effec- It is clear from Figure 4 that there are marked deviations
tively horizontal after about 90 h. It is possible to test from the theoretical curve, and these deviations prove to
directly whether the soluble material being removed had be very informative.
about the same density as amorphous polypropylene. Sup- Comparison of columns 4 and 5 in Table 2 will show
pose that a specimen is soaked as above for 7 days, dried, that for all solvents the significant deviations arise for soak
and a fragment released into the density gradient column temperatures in excess of 50°C. It appears that up to ~4
at the same time as an unsoaked fragment of identical wt% of the soluble material has a density like that of amor-
shape and size. The pre-soaked specimen will settle fairly phous polypropylene, as required by the theoretical curve.
quickly to its final density since there is no more material Some of the 4% is evidently isotactic yet non-crystalline.
to be dissolved, but the other will behave differently. For weight losses between 4 and 12% the points lie to the
According to Figure 3 about 0.68 wt% will go into solu- right of the curve, which is evidence for a structural reor-
tion in the first 7 h, assuming that the mixture of column ganization process during the soaking, the result being a
liquids behaves like pure p-xylene. (This is a reasonable density increase over and above that caused by dissolution
assumption on two counts: the column liquid adjacent to of low density material. At very high temperatures and
the specimen after 7 h is 83% p-xylene by volume, and when the weight loss exceeds 12%, the points start to fall
separate experiments have shown that the weight loss after above the curve and another new feature requires
7 days is the same in pure chlorobenzene as in pure p- explanation.
xylene.) In the period between 7 h and 7 days after inser- To interpret Figure 4 it is necessary to recall certain
tion of the test fragments into the column, the unsoaked morphological features of semi-crystalline polymers. Dur-
piece will still be losing atactic or low crystallinity isotactic ing the crystallization of polyethylene from the melt there
material and will continue to fall as its density increases. is some segregation of low molecular weight material 21~2
The above argument was tested by cutting 4 samples, and this tends to persist in part in uncrystallized form lo-
2 mm square, from film previously soaked in p-xylene at cated within and between the spherulites. Hofmann and
25°C for 7 days, then dried, and 4 similar samples were cut Lauritzen 23 have shown that the disappearance of crystal-
from unsoaked film. Figure 3 predicts that unsoaked speci- linity at temperatures well below the melting point 24-26
mens should lose about 0.44% in weight between 7 h and may be attributed to the early melting of the smallest
7 days, and equation (1) permits the calculation of the cor- crystallites. Baddour et al.27 have pointed out that such
responding increase in overall crystallinity. A crystallinity effects are accentuated by the presence of solvent acting
change of 0.30% corresponds to a displacement of 3.7 mm as a diluent to reduce the melting points of crystallites of
in the density gradient column, which compares well with any size below their 'dry' value. If the smaller crystallites
the observed value of 3.9 -+ 0.5 mm. The implication is that also tend to contain more of the lower molecular weight

390 POLYMER, 1 9 7 6 , V o l 17, M a y


Dissolution o f polypropylene in organic solvents (1): D. A. Blackadder and G. J. Le Poidevin

Table 2 Effect of soak temperature on the density of poly- process (c) becomes less important in absolute terms, as
propylene the following observations confirm.
Consider a polypropylene specimen which had been
ds calcu-
lated soaked at 100°C in decahydronaphthalene (well above the
Soak Mea- Mea- from curve in Figure 4) and another soaked in p-xylene at the
tern- sured % sured equation same temperature (below the curve). After removing all
perature weight ds (3) ~d s the imbibed solvents these dried specimens were re-immer-
Solvent (°C) loss (kg/m 3) (kg/m 3) (kg/m 3)
sed in separate tubes containing fresh p-xylene at 25°C.
p-Xylene 50 2.20 909.0 909.3 --0.3 The results are shown in Table 3. The sample pre-treated
Cumene 50 1.62 908.8 908.9 --0.1 in p-xylene is seen to have absorbed more than twice as
Chlorobenzene 50 1.92 909.0 909.1 --0.1 much of this same solvent at 25°C as the sample pre-
Decahydro- 50 2.22 909.0 909.3 --0.3
naphthalene
treated'in decahydronaphthalene. Even on the basis of
C* (volume of liquid sorbed per unit volume amorphous
p-Xylene 75 4.12 911.3 910.4 0.9
Cumene 75 3.80 910.1 910.2 -0.1
polymer) the uptake at 25°C is much higher for the sam-
Decahydro- 75 4.83 911.6 910,8 0.8 ple pre-treated in p-xylene. This is the opposite of what
naphthalene is generally found for specimens not subjected to pre-
Cumene 95 4.59 912.1 910.6 1.5 treatment, as it will be shown later that C* normally in-
Decahydro- 95 7.68 913.7 912.5 1.2 creases with the density of dry but annealed specimens.
naphthalene The decrease in C* with increased density cannot be ex-
n-Octane 95 6.27 912.2 911.5 0.7 plained in terms of void formation caused by the loss of
p-Xylene 100 8.88 914.6 913.2 1.4 soluble polymer, since the percentage loss is considerably
Cumene 100 6.87 913.8 912.0 1.8 higher for decahydronaphthalene pre-treatment, yet the
Decahydro- 100 29.4 920.8 931.8 -11.0
naphthalene amount ofp-xylene finally imbibed at 25°C is lower.
n-Octane 100 6.00 913.7 911.5 2.2 A possible explanation is that the interlamellar tie-
I-Chloro- 100 3.68 910.5 910.0 0.5 lengths in the two specimens are different after pre-treat-
naphthalene ment, and these lengths determine the extent to which re-
Cumene 107 19.22 916.0 920.8 --4.8 swelling can occur. The specimen pre-treated in decahydro-
n-Octane 107 9.29 914.9 913.6 1.3 naphthalene would have to have the shorter tie-molecules
Cumene 110 16.34 916.3 918.5 -2.2 and this could arise if some of the inter-lamellar chains
n-Octane 110 14.06 916.1 916.9 -0.8 originally present were incorporated into the crystalline
lattice structure as a result of soaking in this good solvent.
Re-swelling at 25°C would then give a low result.
material, then the deviations from the curve in Figure 4 Blackadder and Keniry 8 have shown that if polyethylene
may be interpreted in terms of three parallel processes: film is soaked in liquid p-xylene at 50°C then dried, it
(a) removal of amorphous material having the same den- absorbs a greater amount ofp-xylene vapour at 30°C than
sity as amorphous polypropylene; (b) melting out and
dissolution of small crystallites or crystallites containing
notable amounts of low molecular weight polymer; (c)
inclusion of material described in (a) and (b) into existing Table 3 Effect of solvent pretreatment at high temperatures on
crystallites with a consequent increase in overall density, subsequent reswelling at 25°C in p-xylene
or possibly the creation of new crystallites large enough to
Reswelling in p-
resist solvent. Solvent pre-treatment xylene at 25°C
Process (a), on which the theoretical curve in Figure 4
is exclusively based, evidently covers the lower tempera- Equili- Equili-
tures of soaking and weight losses up to ~4%. Suppose brium Percen- brium
Temper- swelling tage swelling
that at temperatures of 75°C upwards, the consequences of ature X 104 weight dst X 104
soaking were covered by processes (a) and (b) only. Al- (°C) Solvent (m3/kg) loss (kg/m 3) (m3/kg) C*
though m would continue to rise with ds, the points would
lie above the curve, because the dissolved material would 100 p-Xylene 7.0 8.88 914.6 5.53 1.92
100 Decahy- 12.7 29.4 920.8 2.46 1.05
include some polymer from the crystalline regions. The
dronaph-
density of the soaked polymer would therefore be lower thalene
~han if all the soluble material had come from non-crystal-
line regions. (Indeed if a high enough proportion were t d u for both specimens was 908.0 kg/m 3
leached from the crystalline regions it would even be pos-
sible for d s to fall with increasing m, though this was not
Table 4 Effect of liquid pre-treatment on percentage weight loss
observed.) Figure 4 shows that the points concerned ac- of soluble material
tually lie below the curve, and this positively requires the
operation of process (c) in addition to (a) and (b) for Pre-treatment Percentage weight loss
values o f m from 4 to 12%. The process is presumably simi-
Temper- True % Calcu- Calcula-
lar to that induced in pure bulk crystallized polymer on ature Time weight lated ted from
dry annealing at elevated temperatures, but with solvent Solvent (°C) (Days) loss from d s swelling
present, the reorganization occurs at much lower tempera-
tures. At temperatures in excess of 100°C it can be seen p-Xylene 75 7 4.12 5.70 5.0
that the experimental points lie well above the curve on p-Xylene 100 7 8.88 10.8 24.6
Decahydro- 100 7 29.4 19.5 8.15
Figure 4, and this may be interpreted in terms of the in- naphthalene
creasing dominance of process (b). It does not mean that

P O L Y M E R , 1976, V o l 17, M a y 391


Dissolution of polypropylene in organic so~vents (1): D. A. Blackadder and G. J. Le Poidevin
I'0 p-xylene, can be explained as follows. The fact that the
true weight loss is less than the value calculated from ds
points to some incorporation of previously non-crystalline
material into crystallites, while the much larger loss calcu-
lated from the swelfing data points to a considerable length-
0'8 ening of the average interlammellar tie length above the un-
treated value. Blackadder and Keniry 8 showed that sorp-
tion of vapour at quite low activity and temperature by
polyethylene film resulted in a permanent increase in the
average length of interlamellar ties. This can arise because
0"6 bulk crystallized polymer has a wide distribution of tie-
lengths. On the way to sorption equilibriur~ some of the
shorter ties might be pulled out or even fractured, and this
process can hardly be reversed on removing the solvent,
thus resulting in a permanent increase in the average length
of the interlamellar ties. The sorption of liquid p-xylene
0-4
by polypropylene at 100°C has a more drastic effect, as
seen by the very large difference between the measured
weight loss and that calculated from swelling data.
The situation when poly]~ropylene is pre-treated in de-
cahydronaphthalene at 100~C is quite different. The much
02 lower percentage loss calculated from swelling measure-
ments is presumably the result of some recrystallization,
yet at 100°C untreated polymer absorbs almost twice as
much decahydronaphthalene as p-xylene. This fits perfectly
with the expectation that the former solvent would pene-
I I I t I I I trate the polymer to a greater extent than the latter and
0 2 4 6 recrystallization would be easier under the conditions of
t l/2[Ix l(~S l secl/2rn -I )
the pre-treatment.
Figure 5 Reduced sorption curves for polypropvlene and liquid
p-xylene at 25°C. Initial densities: [3, 907.7; O, 914.1; X, 917.9
kg/rn 3 Sorption kinetics
Studies of sorption kinetics are likely to be relevant to
polymer dissolution. Unlike microcrystalline substances,
does untreated film, whatever the vapour activity. Indeed
amorphous polymers swell prior to dissolution 2a, and it will
by extrapolating the relevant isotherms to unit activity it
emerged that the increased volumetric uptake exactly com- be shown in Part 2 that polypropylene also swells. One of
the factors determining the rate of dissolution might thus
pensated for the volumetric loss of soluble material in the
be the rate of movement of the diffusing solvent front as it
pre-treatment. Experiments along similar lines would
penetrates the non-crystalline material before affecting the
appear to offer another method of establishing when the
crystallites.
dissolution of soluble material from polypropylene is
For a plane sheet of polymer, thickness 1, immersed in
accompanied by reorganization of the crystalline and non-
a liquid, it is a standard result that:
crystalline regions.
For this work liquid uptakes were measured rather than
0.0419
vapour isotherms. Specimens of polypropylene film of D=- - (4)
density 908.0 kg/m 3 were immersed in p-xylene (at 75 ° or (tvJl 2)
100°C) or decahydronaphthalene (at 100°C). After 7 days
the specimens were removed from their respective liquids, where D is a diffusion coefficient and tlA is the time re-
dried, reweighed and the losses determined. They were then quired for the amount of liquid imbibed to reach half of
re-immersed in separate 50 cm 3 quantities of p-xylene at its equilibrium value. The intrinsic error in this approxi-
25°C. Unsoaked control specimens were also treated in this mate formula is negligible, but a more serious objection is
way. After 24 h the equilibrium uptakes were determined that it assumes a constant diffusion coefficient. In fact D
and, without exception, samples which had been pre-treated will depend on concentration (among other things) and the
at temperatures above 25°C imbibed more solvent than did use of equation (3) will generate some sort of average value
untreated specimens. The increased liquid uptake was used for the range of concentrations present in the film during
to estimate the equivalent weight lost during the pre-treat- the measurements.
ment on the assumption that the soluble material had a den- Reduced plots for the sorption of liquid p-x_ylene into
sity of about 860 kg/m 3 at 25°C. The results appear in polypropylene films of different densities are presented in
Table 4, together with the true weight losses and values cal- Figure 5, which refers to 25°C. Mt/M= is expressed as a
culated from d s. function of t/l 2, M t being the uptake at time t, and Moo is
For the samples pre-treated in p-xylene at 75°C it is the final equilibrium uptake. The curves have sigmoid
noteworthy that the weight losses arrived at by all three shapes as observed by Rogers et al. 29 for the sorption of
methods are much more comparable than for the other pre- organic vapour into polyethylene at high vapour activities.
treatment procedures. This suggests that for this 75°C Figure 5 shows that liquid is imbibed more rapidly for the
treatment the material removed is largely from non-crystal- higher densities, and Table 5 indicates that there is a re-
line regions and any induced reorganization is small. The duced equilibrium uptake at the higher densities. Note that
wide spread in the loss values for treatment at 100°C in appropriate film densities were obtained by annealing, and

392 POLYMER, 1976, Vol 17, May


Dissolution of polypropylene in organic solvents (1): D. A. Blackadder and G. J. Le Poidevin
Table 5 Effect of density on diffusion coefficient and equilibrium CONCLUSIONS
uptake of p-xylene at 25°C
(1) The Hildebrand solubility paran~eter for polypropylene
Anneal is 16.7 (M J/m3) 1/2. This quantity is obtainable by plotting
temperature Density D X 1012 Equilibrium the swelling caused by various solvents against calculated
(°C) (kg/m3) (m2/sec) uptake (w/w%)
values of the parameter for these solvents and noting the
120 907.7 0.75 10.70 position of the maximum. A correction must be applied
150 914.1 1.08 9.57 to values for n-alkanes obtained from the cohesive energy
163 918.1 1.28 9.05 density. When suitably corrected the solubility parameters
correlate well with the amount of polypropylene dissolved
by each solvent at 50°C and with the amount of solvent
imbibed by the polymer.
the temperatures concerned appear in Table 5. It has been (2) At 25°C the density of polypropylene films cannot be
shown by McCrum a°, Michaels et al. al and by Keniry 32 that correlated with the amount of polymer dissolved by soak-
ing at that temperature. This suggests that the material
for bulk polymer, D increases with annealing temperature
for experiments with low activity solvent vapours and poly- soluble at 25°C is not removed from the crystalline regions,
ethylene. In the present work on polypropylene and where a view for which other evidence is available. It is noteworthy
the solvent activity is unity (pure liquid), it is again evident that polypropylene samples immersed in a density gradient
that the value of D increases with the annealing temperature. column suffer from loss of soluble material, and this source
Michaels et al. 31'33 have expressed the diffusion coeffici- of error should be allowed for when necessary.
cient of a penetrant in a semi-crystalline polymer in the (3) Soaking of polypropylene film in solvents at tem-
form: peratures from 25 ° to 75°C involves dissolution of soluble
material from the non-crystalline regions only, at least for
D = D*/r3 a polymer density of 908 kg/m 3. Up to 4 wt% of the sam-
ples is thought to be a mixture of atactic material and iso-
where D* refers to the same penetrant in wholly amorphous tactic material, all excluded from crystallites. At tempera-
polymer, r is a tortuosity factor to allow for the increased tures from 75 ° to 100°C solvent soaking brings about con-
path length when diffusion occurs round crystallites, and/3 siderable reorganization of the general morphology, pre-
is a chain immobilization factor to account for the restrain- viously non-crystalline chains or parts thereof being incor-
ing effect of crystallites on segmental mobility. It is known porated into new or existing crystallites. In the high tem-
that annealing leads to an increase in fold length which, in perature range of 100 ° to 110°C (just bordering on that
turn, causes a decrease in the width to thickness ratio of
crystalline lamellae and hence a decrease in tortuosity (for
a given volume fraction of obstructive material tortuosity
is increased least by spheres and most by rods). This would
appear to explain the increase in D with annealing tempera-
tures noted here.
Figure 6 shows the sorption data plotted on the basis of
o4[ / ~ x ~ ) ~

C*, the volume of liquid sorbed per unit volume of non-


crystalline polymer (assuming the two phase model), rather
than Mr/Moo, and it can be seen that C* increases with den-
sity. The explanation of this interesting observation lies in
a comparison of the morphology of semi-crystalline poly- O.3
mer with that of crosslinked amorphous polymer. The lat-
ter swells much less than the unlinked form of the same
polymer because the links restrain the uncoiling of chains L
which accompanies sorption. In polypropylene the non-
crystalline regions might thereby be expected to have a
lower sorptive capacity than wholly amorphous polypropy-
/x
lene, volume for volume. Ochiai et al. 34 measured the equili-
brium uptake of two liquids by polypropylene film and
found that C* increased with density as in the present work.
Their explanation virtually constituted a denial of the sim-
ple two-phase model, and the suggestion made by Keniry in
the context of polyethylene 32 appears more convincing. It
hinges on the observation that mats of agglomerated single O-I
crystals swell more than bulk crystallized polymer, the
obvious inference being that tie-molecules inhibit the swell-
ing of the latter and are absent in single crystal agglomer-
ates. Furthermore, the sorptive capacity of unit volume of
non-crystalline material was highest for single crystals, and
bulk polymer tended towards this only as the density was I I i / I
increased by annealing. An increase in the fraction of non- 2 4
crystalline material in the form of chain folds rather than tl/2l-lx idS(secV2rfil)
miscellaneous tangle would appear to favour high Figure 6 Sorption curves for polypropylene and liquid p-xylene
absorption. at25°C. Initial densities: [3,907.7;©,914.1;X,917.9kg/m 3

P O L Y M E R , 1976, Vol 17, May 393


Dissolution of polypropylene in organic solvents (I): D. A. Blackadder and G. J. Le Poidevin
required for total dissolution) the reorganization is over- 10 Natta, G. J. Polym. ScL 1959,34,531
shadowed by the destruction and dis"solution of the smaller 11 Natta, G., Pasquon, I., Zambelli, A. and Gatti, G. Makro-
tool Chem. 1964, 70, 191
crystallites. It appears that there are irreversible changes 12 Blackadder,D. A. and Vincent, P. 1. Polymer 1974, 15, 2
in the average lengths of the interlamellar ties. 13 Blackadder,D. A. and Keniry, J. S. Makromol. Chem. 1971,
(4) Sorption kinetics show that at 25°C liquid p-xylene 141, 211
swells polypropylene most rapidly at high sample density, 14 Danusso, F., Moraglio, G. and Natta, G. Ind. Plast. Mod.
1958, 10, 40
and in addition more liquid is then imbibed per unit volume 15 Hildebrand, J. H. and Scott, R. L. 'Solubility of Non-elec-
of non-crystalline material. This is explained in terms of trolytes', Reinhold, New York, 1950
morphology. 16 Michaels,A. S., Vieth, W. R. and Alcalay, H. H. J. AppL
(5) The results of this paper form the foundation for the Polym. Sci. 1968, 12, 1621
investigation of the conditons under which total, and not 17 Cowie,J. M. G., Ranson, R. J. and Burchard, W. Br. Polym.
J. 1969, 1,187
merely fractional, dissolution of polypropylene occurs. In 18 Brandrup, J. and Immergut, E. I. 'Polymer Handbook',
particular the present results relate to the induction period Interscience, New York, 1966
preceding steady state dissolution. 19 Hildebrand, J. H., Prausnitz, J. M. and Scott, R. L. 'Regular
and Related Solutions', Reinhold, New York, 1970
ACKNOWLEDGEMENTS 20 Hildebrand, J. H. J. Chem. Phys. 1950, 18, 1337
21 Keith, H. D. and Padden, F. J. J. Polym. ScL 1961, 51, $4
One of the authors (G. J. Le P.) is indebted to the Science 22 Lindenmeyer, P. H. Science 1965, 147, 1256
23 Hofmann, J. D. and Lauritzen, J. I. J. Res. Nat. Bur. Stand.
Research Council for a Studentship. (A) 1961, 65,297
24 Turner-Jones, A., Aizlewood, J. M. and Beckett, D. R.
REFERENCES Makromol. Chem. 1964, 75,134
25 Farrow, G. Polymer 1963, 4,191
1 Ueberreiter, K. and Asmussen, F. J. Polym. Sci. 1957, 23, 75 26 Wilkinson, R. W. and Dole, M. J. Polym. ScL 1962, 58, 1089
2 Ueberreiter, K. and Asmussen, F. MakromoL Chem. 1961, 27 Baddour, R. F., Michaels, A. S., Bixler, H. J., de Filippi, R. P.
43,324 and Barrie, J. A. J. AppL Polym. ScL 1964, 8,897
3 Ueberreiter, K. and Asmussen, F. J. Polym. Sci. 1962, 57, 28 Ueberreiter, K. in 'Diffusion in Polymers' (Eds J. Crank, and
187 G. S. Park), Academic Press, London, 1968
4 Asmussen,F. and Ueberreiter, K. J. Polym. Sci. 1962, 57, 29 Rogers,C. E., Stannett, V. and Szwarc, M. J. Polym. ScL
199 1960, 45, 61
5 Asmussen,F. and Ueberreiter, K. Makromol. Chem. 1962, 30 McCrum,N. G. Polymer 1964, 5, 319
52,164 31 Michaels,A. S., Bixler, H. J. and Fein, H. L. J. AppL Phys.
6 Asmussen,F. and Ueberreiter, K. KolloM-Z. 1962, 185, 1 1964, 35, 3165
7 Asmussen, F. and Ueberreiter, K. Kolloid-Z. 1968, 223, 6 32 Keniry, J. S. PhD Thesis University of Cambridge (1971)
8 Blackadder,D. A. and Keniry, J. S. J. AppL Polym. Sci. 33 Michaels,A.S. and Bixler, H. J. J. Polym. Sci. 1961,50,
1972, 16, 1261 413
9 Natta, G., Mazzanti, G., Crespi, G. and Moraglio, G. Chim. 34 Ochiai,H., Gekko, K. and Yamamura, H. J. Polym. ScL
Ind. Milan 1957, 39,275 (,4-2) 1971,9, 1629

394 POLYMER, 1976, Vol 17, May

You might also like