Está en la página 1de 13

From the Chemistry Department, The Manchester College of Science and Technology,

England

The Polymerization of Vinyl Acetate in 14queous Solution


Initiated by Potassium Persulphate at 6OOC.
By A. S. DUNNand P. A. TAYLOR

(Eingegangen am 4. Juli 1964)

SUMMARY:
Polyvinyl acetate precipitates from aqueous solution as it is formed. When potassium
persulphate is used as initiator the polymer suspension is stable and constitutes the prin-
cipal locus of reaction. The effect of reaction conditions on the rate of polymerization and
on the number of particles formed ha5 been studied. The surface charge density appears to
be the factor which determines the number of particles formed and which limits the rate
of reaction.

ZUSAMMENFASSUNG:
Polyvinylacetat f a t bei seiner Bildung aus wainriger Losung aus. Bei Verwendung von
Kaliumpersulfat a l s Initiator ist die Polymersuspension bestandig und stellt den haupt-
sachlichen Ort der Reaktion dar. Der Eidul3 der Reaktionsbedingungen auf die Poly-
merisationsgeschwindigkeit und auf die Zahl der sich bildenden Polymerteilchen wird
untersucht. Es scheint, dal3 die Oberflachedadungsdichte die Zahl der Polymerteilchen be-
stimmt und auch die Reaktionsgeschwindigkeit begenzt.

Introduction

The theory of emulsion polymerization formulated qualitatively by


HARK INS^) and quantitatively by SMITH and E W A R Tis
~ )now generally
accepted. The quantitative theory has been refined by S T O C K M A Y E R 3 ) .
The discrepancies remaining between the theory and the earlier experi-
mental tests of it with styrene by S M I T H 4 ) have been removed by sub-
sequent careful experimental work by BARTHOLOM~ and GERRENS 5, and
by BURNETT and LEHRLE~).
However results with other monomers do not fit the SMITH-EWART
theory in all respects: the dependence of the number of particles formed
(and hence the rate of polymerization) on the emulsifier concentration
diminishes below the three-fifths power predicted by the SMITH-EWART
theory as the solubility of the monomer in water increases ').

207
A. S. DUNNand P. A. TAYLOR

OKAMURAand MOTOYAMA~) have recently shown conclusively t h a t this


deviation from t h e SMITH-EWARTequation arises from the solubility of
the monomer in the aqueous phase. Polyvinyl alcohol is commonly used
industrially as the principal emulsifying agent in t h e polymerization of
vinyl acetate though i t is more typical of t h e agents used t o stabilize
monomer droplets in the suspension polymerization process. Though i t
has recently been shown t h a t critical micelle concentrations can be de-
termined for some monodisperse non-ionic surface active agents 9), poly-
vinyl alcohol is not known t o form micelles lo).HARKINS'S theory identifies
the monomer solubilized in the micelles as t h e locus in which the reaction
is initiated : in the absence of micelles, polymerization might be initiated
in the aqueous phase when the solubility of the monomer in water is
sufficient. Thus a study of t h e polymerization of aqueous solutions of
vinyl acetate should provide information which would clarify the details
of the emulsion polymerization of this monomer. Similar studies have
been made of the polymerization of aqueous solutions of methyl meth-
acrylatell) and acrylonitrile12) but no work on aqueous solutions of vinyl
acetate had been published until t h e work of NAPPERand PARTS13)
appeared recently b y which time the work described in the present paper
had been substantially completed. Several studies on the emulsion poly-
merization of vinyl acetate have been published

Experimental

Materials. Vinyl acetate monomer provided by Vinyl Products Ltd. was redistilled
before use. Potassium persulphate was recrystallized from water: a 1yo solution was pre-
pared just before use. Sodium dodecyl sulphate was purified by HARROLD'S method15):
the critical micelle concentration, determined with a DE NOUYtensiometer was sharp a t
4. M: Polyvinyl alcohol was DuPont "Elvanol" 52-22 containing about 12 mole-%
residual acetate.
Polymerizations were followed dilatometrically using dilatometers with a capacity of
about 21 ml. having a stem approximately 1.7 mm. in diameter. All experiments were a t
60 "C. In some experiments solutions were degassed by repeated freezing and evacuation;
usually, the amount of oxygen in solutions was limited by boiling out the water just before
use.
The soZuGiZity was determined at 20 and 50 "C.to be 2.5 g. vinyl acetate per 100 g. water:
variation with temperature in this range was negligible.
Bromometric titrations. A bromate-bromide solution was acidified to give a solution
0.1222 N in bromine. A reaction time of about one hr. a t 0 "C. was allowed. The method was
tested by determining the purity of redistilled vinyl acetate which was found to be 99.6%.
Determinations I J ~Numbers of Particles. Particle sizes were determined by the light
scattering technique of BURNETT,LEHRLE,OVENALL,and P E A K E R ~ ~ ) a BRICE-
using
PHOENIX Light Scattering Photometer.

208
The Polymerization of Vinyl Acetate Initiated by Potassium Persulphate

Some results were confirmed by electron microscopy of gold shadowed carbon replicas
u3ing a Philips EM 75 electron microscope.

Results
Contraction on polymerization
I n order to convert observed dilatometer contractions to rates of
polymerization, the contraction corresponding t o complete polymeriza-
tion must be known. The polymer formed corresponding t o an observed
contraction was weighed after water and residual monomer had been
removed. Three distinct techniques of removing the liquid residue were
tried : evaporation at 100 "C. after addition of hydroquinone, evaporation
a t 60°C. in a stream of oxygen, and sublimation of the frozen solution.
These gave concordant results, the mean of fifteen determinations being
15.7 f 0.4 yo contraction for complete polymerization. This agrees with
NAPPERand PARTS13) who also find 15.7% for polymerization of mono-
mer in aqueous solution at 60°C. but differs from STARKWEATHER and
TAYLOR'S value of 26.8 yo for bulk monomer a t 70 "C. This might result
from solvation of the vinyl acetate in aqueous solution.

Effect of oxygen
The rate of polymerization increases with time initially (Fig. 1).Since
oxygen inhibits the polymerization such an acceleration could occur
whilst the last traces of oxygen are being consumed. However degassing
the solution rigorously eliminated the inhibition period without affecting
the subsequent course of the reaction. Rigorous degassing made it difficult
to observe the start of the reaction which was masked by expansion as the

Time(min.)

Fig. 1. Effect of oxygen. A Rigorously degassed, maximum rate 3.24y0/min. o Oxygen


content limited by use of boiled out water: maximum rate 3.18y0/min. Vinyl acetate con-
centration 1.0 yo v./v., persulphate concentration 0.01 Yo w./v.

209
A. S. DUNNand P. A. TAYLOR

dilatometer attained thermostat temperature : a short inhibition period


was therefore convenient. I n most experiments the content of dissolved
oxygen was reduced but not eliminated. Fig. 1shows that the rate remains
practically constant at its maximum value between 20 and 45 yopolymeri-
zation. This maximum rate was the same whether the mixtures were de-
gassed or not but rather better reproducibility was achieved with solu-
tions which were not degassed.
Course of the reaction
The course of the reaction can be considered in three stages (i) initial
acceleratory period (ii) period of maximum rate (iii) final deceleratory
period.

Fig. 2. Polymerization of 2 yo v. /v. vinyl acetate with 0.02 yow. /v. persulphate. A Ionic
strength increased by addition of 4% w./v. potassium sulphate. o Ferric-hydrogen per-
oxide initiation. (H,O,) 0.047 M , (Fe+++) 3.5.10-5 M, giving a similar rate of initiation
to 0.02 yo persulphate

Polyvinyl acetate is insoluble in water: the start of the polymerization


is marked by opalescence of the solution. The suspension of polymer
produced is very stable: it does not settle unless centrifuged a t high speed.
Obviously the reaction must initially be a solution polymerization but it
is possible that polymerization will also occur in the swollen polymer as
soon as the second phase is formed. That the polymer phase soon becomes
the more important locus of the reaction is clear from Fig. 2 where the

210
The Polymerization of Vinyl Acetate Initiated by Potassium Persulphate

rate of polymerization of a solution of vinyl acetate initiated b y potassium


persulphate is compared with the rate of polymerization of a similar solu-
tion initiated b y t h e hydrogen peroxide-ferric ion system. The concen-
trations of initiators were chosen t o give similar rates of initiation, but t h e
later stages of the polymerization are very much slower with the peroxide-
ferric initiator : there is no acceleratory period, the rate decreases and
t h e reaction practically ceases a t less t h a n 100 yo polymerization. The
precipitated polymer does not form a stable sol b u t settles out as a swollen
gel. Monomer dissolved in t h e gel phase is removed from t h e locus of t h e
reaction and is not polymerized. Polymerizations photoinitiated with
uranyl nitrate as photosensitizer behave similarly. With persulphate
initiation solution polymerization gives place t o emulsion polymerization
with t h e swollen polymer particles as the principal locus of reaction with
the difference t h a t no monomer droplets are present a t any stage.
Though t h e rate of reaction is not strictly constant a t any time there
is a considerable period (20 t o 4 5 % polymerization in Fig. 1) when t h e
rate is little different from its maximum value. Since no method is known
b y which a rate constant which will characterize t h e whole course of the
reaction can be derived, this maximum rate seems t o be the most useful
characteristic t o discuss with regard t o the effect on t h e rate of variation
in the conditions of reaction. NAPPERand P A R T Shave ~ ~ ) shown that, on
the assumption t h a t the rate constants do not depend on conversion b u t
t h a t the initial acceleration is t o be attributed t o increasing numbers of
radicals per particle, a n expression can be derived which shows t h a t the
square root of t h e rate of polymerization should be proportional t o the
time. Such a relation is followed during the acceleratory period in ex-
periments in which surface active agents were not added but not in those
in which they were present. Rates observed during the third stage in the
absence of surface active agents fit a first order equation but this relation
also breaks down when surface active agents are present.

Distribution of monomer between phases


Vinyl acetate is soluble in both the aqueous and the polymer phase so
t h a t monomer is extracted from t h e water b y the polymer as i t is formed.
The distribution of monomer between phases was determined by allowing
a known volume of monomer t o come t o equilibrium a t about 20 "C. with
a measured amount of completely polymerized polyvinyl acetate latex.
Mixtures were then centrifuged a t 16,000 t o 18,000 r.p.m. a t 5°C. in
tubes covered t o prevent evaporation of monomer. Monomer remaining
in the aqueous phase was determined bromometrically.

211
A. S. DUNNand P. A. TAYLOR

If Mp is the percentage by weight of monomer in the polymer phase


and M, is the percentage by weight of monomer in the aqueous phase,
then
Mp = 13.7 Mw2 (1)
NAPPERand PARTS13)obtain a similar relationship. The reason for the
divergence from the simple NERNST Distribution Law might be sought in
the dependences on concentration of the activity coefficients of vinyl
acetate in the two phases.
Eflect of stirring
Stirring has been shown to affect the rate of polymerization of aqueous
solutions of acrylonitrile 12). A Camlab Ltd. underwater magnetic stirrer
was arranged under a dilatometer with a flat base but stirring made no
difference either to the rate of polymerization or to the particle size of
t h e fully polymerized latex.
Rate of initiation
The rate constantls) for the thermal decomposition of persulphate a t
60 "C. is 4.07.10-4 min-l whence the rate of initiation should be 6.0.10-7
mole 1-1 min-l a t a concentration of 7.32 M. Polymerization could
be inhibited by hydroquinone or p-benzoquinone :the rate of consumption
of hydroquinone was 3.6 mole 1-1 min-l and that of p-benzoquinone
1.8.10-7 mole 1-1 min-l which imply the same rate of initiation if i t is
assumed that hydroquinone reacts with one and benzoquinone with two
sulphate radicals. This would be consistent with a 6 0 % efficiency of
initiation.

InJluence of initiator and initial monomer concentration on the rate of


polymerization and the size of particles formed
At a fixed initial monomer concentration (Table 1) the maximum rate
of polymerization was found t o be proportional to the initiator concen-
tration t o the power 0.64. At a fixed initiator concentration (Table 2) the
maximum rate increased with initial monomer concentration. The size
of the particles in the fully polymerized latex decreased with increasing
initiator concentration but their number increased (Table 1). The particle
size increased with increasing initial monomer concentration (Table 2) :
t h e number of particles formed decreases with increasing monomer con-
centration but is almost constant when the initial monomer concentration
is greater than 1 yov./v.

2 12
The Polymerization of Vinyl Acetate Initiated by Potassium Persulphate

Table 1. Variation of rate and number of particles with initiator concentration. Initial
monomer concentration 1yo v. /v.

Particle diameter
% K2S20, Maximum rate No. particles/mL
w./v. D w (4 (. 10-12)
(at lo0 yo conversion)
I I I
I I I
0.005 1.9 2950 0.74
0.010 3.2 2740 0.93
0.015 4.0 2600 1.09
0.020 4.8 2480 1.25
0.030 - 2440 1.30
0.040 7.5 2310 1.54

Initial monomer Maximum rate Particle diameter No. particles /ml.


yow./v. yo/min. (. 10-12)
0.1 - 510 15.1
0.2 - 1000 3.83
0.3 2.3 1360 2.29
0.5 3.1 1760 1.74
1.0 4.8 2480 1.25
1.5 4.9 3100 0.96
2.0 4.9 3210 1.15

Initial monomer Initiator YO Particle diameter No. particlesIml.


yov./v. Yo w. IV. Conversion D w (4 (. 10-12)

5 1400 1.38
16 1700 1.48
37 2180 1.17
62 2900 0.85

1.0 1 0.01
100
6
17
3280
740
1250
0.91
91.0
28.2
32 1590 19.8
49 1780 16.0
66 1980 14.8
100 2280 12.7

213
A. S. DUNNand P. A.TAYLOR

Variation of the number of particles during the course of polymerization


The initial acceleratory period which has been observed could be con-
nected with a n increase in t h e number of particles. However determina-
tion of t h e numbers of particles a t various stages during t h e course of a
polymerization (Table 3) showed, on t h e contrary, t h a t the number of
particles decreased with time.

Effect of ionic strength


The effect of the total concentration of electrolyte on the course of t h e
reaction was not investigated in detail, but Fig. 2 shows t h a t t h e addition
of neutral salt has a large effect on the rate. A stable sol was not formed a t
this salt concentration: after 34 min. t h e polymer precipitate began t o -
settle out.
Eflect of sodium dodecyl sulphate
Fig. 3 shows the effect of sodium dodecyl sulphate on t h e rate of re-
action. The maximum rate is increased only slightly but t h e duration of
t h e acceleratory period is much reduced. The addition of small quantities
of the surface active agent produces a marked diminution in t h e size of the
particles formed (Table 4) : above the critical micelle concentration (de-
termined as 0.11 yo w . / ~ .for t h e material used) t h e particle size remains
constant.

Table 4. Effect of sodium dodecyl sulphate (SDS) on particle size (Dw).


Initial monomer concentration 2.0 yo v./v. Potassium persulphate 0.02 Yo w. /v.

(100 yo conversion)

330

Effect of polyvinyl alcohol ,


Polyvinyl alcohol (Table 5) reduces t h e maximum rate: it does not
reduce the acceleratory period. It is less effective in reducing particle size
t h a n sodium dodecyl sulphate.

yo w./v. polyvinyl alcohol . . . 0 0.120 0.240 0.288 0.360


D, (A) ( l O O ~ oconversion) . . 3060 1450 930 - 780
Maximum rate (%/min.) . . . 5.15 4.60 3.54 3.27 2.54

2 14
The Polymerization of Vinyl Acetate Initiated by Potassium Persulphate

Time (min

Fig. 3. Effect of sodium dodecyl sulphate on the rate of polymerization of 2 yo v./v. vinyl
acetate with 0.02% persulphate. yo w./v. sodium dodecyl sulphate: - 0, --- 0.00046,
_ _ _ 0.004-0.049

Discussion

Comparison with polymerizations initiated by hydrogen peroxide or


uranyl nitrate makes i t clear that the reaction initiated by potassium
persulphate takes place mainly in the precipitated phase of swollen poly-
mer. Taking the value19) of kP/kt1l2to be 0.6 a t 60°C. and rate of initi-
ation to be 3.6 - lo-' mole 1-1 min-l the rate of solution polymerization in
a 0.23 M solution of vinyl acetate is calculated as 6.1. mole 1-1 min-l
whereas the maximum rate observed in such a system was 1.1. mole
1-1 min-l, twenty times greater. Whilst the polymerization must start in
solution, the reaction must be considered t o be a polymerization of the
emulsion type for the greater part of its course. I n some experiments the
monomer concentration in the aqueous phase, was measured a t various
stages in the polymerization. The monomer concentration in the polymer
particles can be calculated from these results: it is very low. Even a t the
earliest stage, the monomer concentration in the polymer particles
(Table 6) is comparable with that prevailing during the final stage of a n
ordinary emulsion polymerization. BENGOUGH and MELVILLE20) have
shown that there are large variations in the values of rate constants in
such viscous media. If it is assumed that there is no more than one grow-
ing radical in a particle then an explanation of the variation of the rate of
reaction with time might be sought in the variation of the propagation
rate constant with the monomer concentration in the polymer particles.
However the reaction differs from an ordinary emulsion system in
that the monomer concentration in both phases decreases continually

215
A. S. DUNNand P. A. TAYLOR

throughout the reaction. Because the amount of the polymer phase


progressively increases, monomer dissolved in the polymer phase passes
through a maximum which is parallel to the maximum in the rate of
reaction (Fig. 4).

% Polymerization

Fig. 4. Correlation of the rate of polymerization - - - with the total amount of monomer
dissolved in the polymer phase -

The size of the suspended particles is some ten times larger than those
formed in the experiments of SMITHand later workers4-6). It has been
shown 21) that several polymer radicals can co-exist in such large particles.
Calculations have been made of the number of radicals per particle which
would be required t o explain the observed maximum rates of reaction,
assuming the propagation rate constant, k,, to be constant and to have
a value19)of 3,000 1. mole-1 sec-l. From the data of Table 3, ii the average
number of radicals per particle was calculated (Table 7) from the observed
number of particles per litre, using M,, the monomer concentration in the
particles derived from the concentration of monomer remaining in the
aqueous phase determined by bromometric titration and the distribution
expression. Then
- dM/dt = fi kp (Mp) N/N* (2 )

where NA is AVOGADRO’s number.


The low value of ii (ii = 0.5 in an ordinary emulsion polymerization)
is found a t a stage when particle cdescence is occurring: under these
circumstances termination when particles coalesce by mutual reaction
of polymer radicals initially in different particles may be important. The
calculated number of active centres attains values which, a t first sight,
appear implausibly high. However it is conceivable that such a large
number of growing radicals could co-exist in a particle which was so

216
The Polymerization of Vinyl Acetate Initiated by Potassium Persulphate

Table 7. Variation of the number of radicals (ii) per particle with conversion

2.0% v./v. monomer, 0.02% w./v. persulphate


yo polymerization . . . . . .
......... I 5
4.4 j 16
3.4
37
2.1
62
0.6
100

Rate. 103 (mole I.-' min-')


n. . . . . . . . . . . . . .
..
1 3.36
1.8
6.26
4.2
10.45
14.2
9.6
61.5

1.0% v./v. monomer, 0.01 yo w./v. persulphate


yo polymerization . . . . . . 6 17 32 49 66
M, molar . . . . . . . . . . 0.33 0.21 ~ 0.08
-
n. . . . . . . . . . . . . . 1.38 2.57 3.65

viscous as t o prevent mutual termination between polymer radicals and


i n which the only alternative termination process involves t h e approach
of a charged sulphate radical t o a particle of similar charge.
The experiments in t h e presence of sodium dodecyl sulphate are re-
markable in t h a t t h e maximum rate of polymerization is little influenced
by changes in t h e number and size of the particles :the number of polymer-
izing radicals in t h e system is evidently not influenced by such changes.
Adsorption of the dodecyl sulphate anions will permit t h e surface
charge density necessary t o render small particles stable and prevent
further coalescence t o be attained a t smaller particle sizes t h a n when the
sulphate end groups are the only charged groups in the system. However
the charge density necessary for stability depends on t h e ionic strength
of t h e system as is shown b y the experiment (Fig. 2) with added salt.
This complicates interpretation of the experiments with sodium dodecyl
sulphate since t h e ionic strength was not maintained constant. The
evidence is t h a t there is a rapid decrease in the number of particles in t h e
system during t h e first acceleratory stage of the reaction. During this
stage, termination between polymer radicals initially in different parti-
cles may be a n important factor reducing the rate of reaction. This is
supported b y t h e effect of sodium dodecyl sulphate in reducing t h e dura-
tion of t h e acceleratory period: the number of primary particles which
must coalesce t o give stable particles is smaller, the opportunity for
termination between radicals in different particles is reduced, so t h a t t h e
maximum rate is attained a t a n earlier time. I n the experiment with
added potassium sulphate coalescence probably continues throughout
resulting in a reduction of the maximum rate attained as well as the
ultimate precipitation of the suspension.
It is known 22) t h a t transfer reactions between vinyl acetate and poly-
vinyl alcohol can occur resulting in t h e formation of graft co-polymer.

217
A. S. DUNNand P. A. TAYLOR

Transfer reactions can retard the rate of polymerization 23) because of the
introduction of a n alternative termination reaction.
Termination between a polyvinyl acetate and a polyvinyl alcohol
radical may be favoured because t h e approach of a charged radical t o a
charged particle is not involved. But polyvinyl alcohol increases the
stability of smaller particles (Table 5): these must have a lower surface
charge density which should increase the rate of termination b y sulphate
radicals.
Thus t h e factors which determine the rate of polymerization and the
size of the particles formed in this reaction are quite different from those
operative in the emulsion polymerization of styrenel-Q : i t is hoped t h a t
application of the theory of t h e stability of lyophobic t o this
system will provide a more quantitative explanation of its behaviour.

We wish t o t h a n k VINYLPRODUCTS LTD. for a grant t o P.A.T. during


the tenure of which this work was carried out.

l) W. D. HARKINS,J. Amer. chern. SOC.69 (1947) 1428.


2, W. V. SMITH and R. H. EWART, J. chem. Physics 16 (1948) 592.
3, W. H. STOCKMAYER, J. Polymer Sci. 24 (1957) 314.
') W. V. SMITH,J. Amer. chem. SOC.70 (1948) 3695.
5, E. BARTHOLOME, H. GERRENS, R. HERBECK,and H. M. WEITZ, 2. Elektrochem., Ber.
Bunsenges. physik. Chem. 60 (1956) 334.
6, G. M. BURNETTand R. S. LEHRLE,Proc. Roy. SOC.[London] A 253 (1959) 331.
') H. CFIERDRON, I.U.P.A.C. Symposium on Macromolecules, Wiesbaden 1959, Kurz-
rnitteilungen 111 C 7.
*) S. OKAMUIU and T. J. MOTOYAMA,J. Polymer Sci. 58 (1962) 221.
*) J. M. CORKILL, J. F. GOODMAN, and R. H. OTTEWILL,Trans. Faraday Soc. 57 (1961)
1627.
la) J. T. O'DONEELL,R. B. MESROBIAN,and A. E. WOODWARD, J . Polymer Sci. 28 (1958)
171.

218
The Polymerization of Vinyl Acetate Initiated by Potassium Persulphate

11) J. H. BAXENDALE, M. G. EVANS,and J. K. KILHAM,Trans. Faraday SOC.42 (1946) 668.


B. ATKINSON and G. R. COTTEN, Trans. Faraday SOC.54 (1958) 877.
12) F. S. DAINTONand P. H. SEAMAN, J. Polymer Sci. 39 (1959) 279; W. M. THOMAS, E. H.
GLEASON, and G. J. MINO, J. Polymer Sci. 24 (1957) 43; D. E. MOOREand A. G. PARTS,
Makromolekulare Chem. 37 (1960) 108.
13) D. H. NAPPERand A. G. PARTS, J. Polymer Sci. 61 (1962) 113.
W. J. PRIEST,J. physic. Chem. 56 (1952) 1077; D. M. FRENCH, J. Polymer Sci. 32
(1958) 395; R. PATSIGA, M. LITT, and V. STANNETT, J. physic. Chem. 64 (1960) 801.
'1 S. P. HARROLD, J. Colloid Sci. 15 (1960) 280.
16) G. M. BURNETT, R. S. LEARLE,D. W. OVENALL, and F. W. PEAKER,J. Polymer Sci. 29
(1958) 417.
17) H. W. STARKWEATHER and G. B. TAYLOR, J. Amer. chem. SOC.52 (1930) 4708.
la) L. GREEN and 0. MASSON,J. chem. SOC.[London] 97 (1910) 2083.

'1 C. H. BAMFORD, W. G. BARB,A. D. JENKINS, and P. F. ONYON,The Kinetics of'Vlny1


Polymerizations by Radical Mechanisms. Butterworth, London 1958, p. 87.
20) W. I. BENGOUGH and H. W. MELVILLE, Proc. Roy. SOC. A 230 (1955) 429.
21) C. P. ROE and P. D. BRASS,J. Polymer Sci. 24 (1957) 401.
22) F. D. HARTLEY, J. Polymer Sci. 34 (1958) 397; C. C. LIN, J. Chin. chern. SOC.(Formosa)
6 (1960) 154 through Chem. Abstr. 55 (1961) 9941h.
23) G. M. BURNETT and L. D. LOAN,Trans. Faraday SOC.51 (1955) 214.
m, B. V. DERJAGUIN and L. D. LANDAU, Acta physicochim. USSR. 14 (1941) 633; E. J. W.
VERWEYand J. T. G. OVERBEEK, Theory of the Stability of Lyophobic Colloids.
Elsevier, -4msterdam 1948.

219

También podría gustarte