You are on page 1of 13

JOURNAL OF POLYMER SCIENCE: Polymer Physics Edition VOL.

12, 1771-1783 (1974)

Temperature Dependence of Viscosity-Shear Rate


Behavior in Undiluted Polystyrene

RICHARD C. PENWELL* and WILLIAM W. GRAESSLEY,


Chemical Engineering Department and Materials Research Center,
Northwestern University, Evanston, Illinois 60201 and ANDRE KOVACS,
Centre de Recherches sur les Macromolecules, Strasbourg, France

Synopsis
The time-temperature superposition principle is well-established for linear visco-
elastic properties of polymer systems. It is generally supposed that the same principle
carries over into nonlinear phenomena, such as the relationship between viscosity 8 and
shear rate +. Guided by this principle and the forms of various molecular theories, one
would expect that 7 -+ data on the same polymer a t different temperatures would
superimpose when plotted as 7/70 versus yao/pT, 80 being the limiting viscosity at low
shear rates, p the polymer density, and T the absolute temperature. Data on poly-
styrene melts, obtained in a plate-cone viscometer, appear systematically to violate this
principle in the range 140-190. Such anomalies are absent in concentrated solutions
of polystyrene. The trends are similar to those reported by Plazek in the steady-state
compliance of polystyrene melts near T,, but they appear to persist to higher tempera-
tures than the compliance anomaly.

INTRODUCTION
The shear rate dependence of viscosity in many polymer systems can be
expressed conveniently in reduced variable form :

The viscosity TJ is the ratio of steady state shear stress u to shear rate ),
is the limiting viscosity at low shear rates, T~ is a characteristic time con-
stant which locates in some suitable manner the vicinity of onset of shear
rate dependence (TJdeparts from TJO in the region where ) = Y o = 1 / ~ ~ ) ,
and V is the viscosity master function. I n concentrated solutions and
melts TJO and TO depend on many system variables (temperature, average
molecular weight, concentration, etc.). The form of V , on the other hand,
appears to be controlled mainly by the molecular-weight distribution of the
polymer. In this paper we report viscosity-shear rate data on polystyrene
melts measured over an extended range of temperatures. These data were
analyzed through eq (1) to obtain values crf T J ~and T ~ . Their variations
* Present address: Xerox Corporation, Webster, N.Y.
1771
@ 1974 by John Wiley & Sons, Inc.
1772 PENWELL, GRAESSLEY, AND KOVACS

with temperature appear to differ in a manner which is contrary to expec-


tations, based on conventional shear rate-temperature superposition
principles.

SHEAR RATE-TEMPERATURE SUPERPOSITION


The relationships between 70 and molecular properties have been re-
viewed recently,' and are particularly w-ell-established in polystyrene sys-
tems. A viscosity time constant r0 can be determined in a number of ways
from experimental but in all cases TO is of the order of magnitude
of the longest relaxation times in the linear viscoelastic relaxation spectrum.
I n a number of molecular theories5f6TO for narrow distribution polymers is
assumed to be proportional to T R ~the longest relaxation time of the Rouse
molecular theory7t8

in which M is the polymer molecular weight, c the polymer concentration


(wt/vol), and RT the Boltzmann factor. I n melts the concentration is
replaced by polymer density p .
Experience has shown that the explicit M and c dependence of TO is some-
what different for concentrated systems than that given by eq. (2).223.9-13
Similar differences are seen in the mean terminal relaxation time rmof the
linear viscoelastic s p e ~ t r u r n . 8 ~Otherwise
~~'~ the theoretical forms for T~
and rmare taken to be correct, such that they are both expected to be pro-
portional to v0 and inversely proportional to pT. Thus, for the same poly-
mer melt a t different temperatures:
7OF
70 = - (3)
PT
in which F , like V , is a property which depends on molecular structure but
is independent of temperature. For r, the form is supported by the well
tested principle of time-temperature superposition for linear viscoelastic
properties.s It seems well founded for amorphous systems as long as the
temperature is not too near T,.15
There is also considerable data supporting eq. (3).10,16-22Based on
eq. (3), reduced curves of 7 / q oversus qoY/pTfor the same sample a t different
temperatures should superimpose. Within experimental error, and con-
sidering the comparatively weak dependence of pT on temperature (in
polystyrene, for example, pT increases by about 9% from 150C to 2OO0C7
while 70 changes by about 160-fold), plots of 9/70 versus 707 should also
superimpose. According to the same principle, logarithmic plots of r]
versus u a t different temperatures should be superimposable by shifts
parallel to the viscosity axis alone. Observations that the temperature
coefficient of viscosity a t constant shear stress is independent of shear
stress is an indirect confirmation of the same superposition principle.
VISCOSITY-SHEAR RATE 1773

SAMPLES AND EXPERIMENTAL PROCEDURES


Four linear and four branched polystyrenes were used in the study. All
were polymerized anionically. Sample 8-109 is one of a series of exten-
sively studied narrow-distribution polystyrenes prepared some years ago
by the Dow Chemical Company. Sample PC-4 is a narrow-distribution
linear polymer supplied by the Pressure Chemical Company. The re-
mainder of the samples were prepared at C.R.M. in Strasbourg by Dr.
J. G. Zilliox. Sample L-1 is a narrow-distribution linear polymer, L-2 is
a broad-distribution linear polymer with AT, similar to L-1, and B-1, B-2,
B-3, and B-4 are narrow-distribution star-branched polymers. Character-
ization data are given in Table I.
The measurements of steady-state viscosity for all samples except PC-4
were carried out with a Kepes plate-cone viscometer. The details of its
design and operation are given elsewhere.23 The cone is the center portion
of a rotating cup. The torque transmitted to the cone is recorded me-
chanically. The heating chamber encloses the system with relatively small
clearances; this and internal compensators provide good temperature con-
trol. The cone used in all measurements with this instrument was 3.2 cm
in diameter and provided a gap angle of 3.5".
The PC-4 data and additional data on S-109 were obtained a t the North-
western laboratories with a modified Weissenberg rheogoniometer (model
R-16). The original torque transducer was replaced by a piezo-electric
load cell as described by B r ~ d k e y . ~
The
~ normal stress detecting system
was replaced by a vertically rigid lower member, the instrument being
operated therefore as a simple plate-cone viscometer. Shear stress and
shear rate were calculated with the standard equations for plate-cone
viscometry .
Figure 1 shows viscosity-shear rate results on PC-4 a t 175C for various
cone angles and diameters. The effect of diameter, which tests the ability
of the instrument to provide the same shear stress for different magnitudes of
steady-state transmitted torque, appears to be negligible. The effect
of gap angle, which we regard as mainly a test for viscous heating or sec-

TABLE I
Molecular Properties of Polystyrene Samples
Sample B,2 Branches/molecule
s-109 162,000 180,000 0
PC-4 386,000 41 1,000 0
L- 1 383,000 404,000 0
L-2 195,000 375,000 0
B1 - 273,000 13.9
B-2 - 1,060,000 9.6
B-3 - 2,050,000 10.7
B-4 - 2,750,000 14.4
* Osmotic pressure measurements.
b Light scattering measurements.
1774 PENWELL, GRAESSLEY, AND KOVACS

10'

Q)
.-m
0
n

I , , 1 3 1 1 I I b I I I l l

0 .I I.o

y sec-'
Fig. 1. Viscosity vs. shear rate at 175OC for linear polystyrene PC-4, showing results
obtained with various plate-cone geometries in the Weissenberg rheogoniometer. Sym-
bols: ( A ) '1 gap angle and 5 cm diameter; (0)2' gap angle and 5 cm diameter; ( 0 )4"
gap angle and 5 cm diameter; ( 0 )2' gap angle and 2.5 em diameter.

ondary flow-S, is fairly small. No evidence of viscous heating was ob-


served. The platen temperature and the shear stress remained constant
even when shearing was prolonged for an hour or more. Gap angle appears
to enter principally in determining the maximum shear rate ymsxfor which
stable steady shear flow can be achieved, as noted in earlier Roughly
speaking ayma,= constant for a given sample, where a is the gap angle.
Most data were gathered with a cone of 5.0 cm diameter and 2" gap angle.
Due apparently to the relatively narrow molecular weight distribution
of the samples and the wide range of shear rates available in the viscometer,
it was possible in all cases to reach limiting behavior and determine 710,
values of which are listed in Table 11. The values of 10 obtained a t the
two laboratories for sample S-109, the only common sample, are compared
in Figure 2.

RESULTS
Figure 3 shows the viscosity-shear rate behavior of linear sample PC-4
from 165C to 217C. Figure 4 shows the same data plotted in the re-
duced form 1 / 9 0 versus q o y / p T . Density at each temperature was evalu-
ated from the following equation:25
1
-
P
= 0.973 + 5.8 X (T - 94C) (4)

It is apparent that superposition is not achieved. Figure 5 shows the


data on S-109 from CRM plotted in the alternative reduced form 7 / 9 0
versus a / p T . Again superposition is not achieved and, as with the PC-4
data, Newtonian behavior persists to higher reduced shear rate or shear
VISCOSITY-SHEAR RATE 1775

TABLE I1
Rheological Parameters of Polystyrene Samples
Labora- vo x lo*,
Sample Za tory r, "C poise n,sec
-- I__. ..-. -
s-109 m CRM 142 3.53 114 112.5 0.99
149.5 1.15 37.7 36.0 0.95
159 0.24 10.3 7.38 0.72
169 0.081 3.88 2.46 0.634
174.5 0.0365 2.33 1.09 0.468
m NU 150 0.720 28.6 22.5 0.787
159 0.202 10.4 6.21 0.597
170 0.0521 3.92 1.58 0.403
200 0.00425 0.480 0.122 0.254
PC-4 0) NU 160 5.22 136 367 2.70
165 2.86 67.8 199 2.93
172 1.38 32.3 94.9 2.93
175 0.895 26.3 61.3 2.33
179 0.627 18.1 42.6 2.35
193.5 0.150 5.71 9.96 1.74
217 0.0359 1.54 2.30 1.49
L-1 m CRM 149 11.9 364 837 2.29
158.5 2.4 86.9 166 1.91
167.5 0.825 30.3 56.3 1.86
174 0.37 21.6 25.0 1.16
G2 5 CRM 149 15.1 580 984 1.70
161 3.00 123 192 1.56
171 0.585 31.7 36.8 1.16
179 0.276 24.4 17.1 0.701
B- 1 5 CRM 130.5 0.76 14.0 30.6 2.18
139.5 0.131 2.11 5.17 2.45
143 0.0735 1.22 2.78 2.28.
B-2 5 CRM 149 3.64 138 670 4.85
159.5 0.92 35.7 166 4.65
167.5 0.31 12.1 55.5 4.58
177 0.107 5.00 18.8 3.76
181 0.0695 4.17 12.2 2.93
B-3 I CRM 170 5.80 290 1998 6.89
181 1.11 76.9 375 4.88
191 0.394 36.4 130 3.57
199 0.18 20.1 59.0 2.94
B-4 1 CRM 185 1.67 100 753 7.53
203 0.400 25.6 175 6.84
* Parameter of theoretical master curve used to obtain values of ro.a6

stress at lower temperatures. The same pattern, and to more or less the
same degree, was found in the other samples as well. As expected, reduced
plots of 7/70 versus 707;and 7/70 versus cr were indistinguishable in appear-
ance from their counterparts (Figures 4 and 5 ) containing the pT factor.
It was found that the viscosity curves could be superimposed if arbitrary
shifts along the shear rate or shear stress axis were allowed. Moreover, the
resulting curves had shapes which were similar to a number of theoretical
1776 PENWELL, GRAESSLEY, AND KOVACS

IC

IC

70
(poise)

IC

10

Fig. 2. Zero-shear viscosity vs. temperature for linear polystyrene S-109. The open
circles (0)are data obtained on the Kepes viscometer; the filled circles ( 0 ) are data
obtained on the Weissenberg rheogoniometer.

master curves calculated for the Zimm-Schulz molecular-weight distribu-


tion.26 The shapes depend on a parameter 2, which is related to poly-
dispersity, according to the theory, by:

For purposes here the curve which best fits all the data for a given polymer
was used, regardless of its distribution, simply as a consistent method for
assigning values of T O . Figures 6 and 7 are typical of the results. The
values of 2 used for each polymer, and the resulting values of ro at each
temperature are given in Table 11. Also included in Table I1 are values
of T R and T E / T O . If the conventional temperature superposition principle
were valid, ~ ~ / wouldr o be the same at all temperatures for each polymer
although varying somewhat from one polymer to another. The systematic
decrease in T R / T O with increasing temperature merely displays in another
way the departures from conventional superposition theory seen in Figures
VISCOSITY-SHEAR RATE 1777

0
0

I I I 1 1 1 1 1 1 I I 1 I I I I I I
0 0.0I 0.1 I.o
Shear Rate y , (sec")

Fig. 3. Viscosity vs. shear rate at various temperatures for linear polystyrene PC-4.

O1 k 100 1000

Fig. 4. Reduced viscosity vs. reduced shear rate for linear polystyrene PC-4. Symbols:
(A) 165C; (0)172C; (0) 175C; ( 0 ) 179C; (m) 193C; (A) 217OC.

4 and 5 : the viscosity-shear rate time constant increases with decreasing


temperature, but less rapidly than qo or q o / p T .
Figures 8 and 9 show r R / r 0 versus T behavior for linear polymers of nar-
row distribution. Aside from differences in magnitude, presumably due
to the differences in molecular weight, the behavior shown in Figures 8 and
9 is very similar. The same trend is present in the branched polymers
also, but is somewhat less definite. Figure 10 shows a comparison of T R / T O
versus T among all linear polymers in the study. Differences in scale
have been eliminated by dividing the % / T O values for each polymer by its
estimated value at 160C. The scatter is substantial, but it appears that
1778 PENWELL, GRAESSLEY, AND KOVACS

I I , , , , / I I

1.0 -

0.3-

0.11 ' " ' ' I '


I I / , , , , I I , , , l l , l

0 0.I 0.3 I .O 3.0 10.0

y Tot2
Fig. 6. Viscosity-shear rate master curve for linear polystyrene S-109. The solid line is
that for 2 = m from Ref. 26. Symbols are the same as in Figure 5.

I
I .o

0.I I I , 8 I ! ,!I I I I , 1 1 1 1

0.I 0.3 I.o 3.0 10.0

y r. 1 2
Fig. 7. Viscosity-shear rate master curve for star-branched polystyrene B-2. The
solid line is that for 2 = 5 from Ref. 26. Symbols: ( 0 ) 149C; (0)159.5OC; ( 0 )
167.5OC; (U) 177C; (A) 181C.
VISCOSITY-SHEAR RATE 1779

3-
0 0

2-

I -

0
I40 I60 180 200 220
TOC

Fig. 8. Ratio of Rouse to experimental relaxation time as a function of temperature for


linear polystyrene PC-4.

a single function, perhaps becoming a constant at high temperatures, would


fit the data. With the approximation of two straight lines:

T > 190C

TR
-
TO
= [1 + 0.02(190 - T ) ] T < 190C

Estimated values of (TR/T0)190 for the linear polymers are somewhat below
those predicted from a recent correlation of T R / T O for concentrated solu-
tions, and melt data obtained in capillary flow.11
I n contrast to the anomalous temperature dependent of T O ,the tempera-
ture coefficient of 90 was unexceptional. Comparisons of 70 values on
sample S-109 with values obtained by other workers is described else-
where.27 The same form for the temperature coefficient UT (a WLF equa-
tion for temperatures below 190C and an activation energy equation a t
higher temperatures) and the published numerical coefficients for poly-
~ t y r e n e ' ~fitted
, ~ * data on UT for both the linear and branched samples.
Viscosities for different samples a t the same temperature differed according
to molecular weight for the linear samples, although those for samples L-1
and L-2 were consistent with molecular weights lower than the reported
values by approximately 17%. Values for the branched polymers in all
cases were substantially less than those of linear molecules of the same
molecuIar weight. We have not attempted a more detailed analysis
based on structure.
1780 PENWELL, GRAESSLEY, AND KOVACS

0
0
0.6-

0
0.4-

o'2
?40 160 180T OC m 220

Fig. 9. Ratio of Rouse to experimental relaxation time as a function of temperature


for linear polystyrene S-109. Symbols: ( 0 )data obtained on Kepes viscometer; (0)
data obtained on Weissenberg rheogoniometer.

DISCUSSION
These results on the temperature dependence of 7 versus 9 appear to be
a t variance with other studies on polystyrene melts.22 In the latter the
conventional superposition principle was confirmed either by direct exami-
nation of q/qo versus TO?; plots or equivalently by the observation that the
activation energy at constant u, E,, is independent of u. However, the
data in these studies were obtained a t high temperatures, generally above
190C, where superposition should be Valid even according to the current
results. I n the only study of temperature effects on 7 versus 9 behavior
which extended appreciably below 190C, the data a t lower temperatures
were not extensive enough to show up deviations.'7 Measurements of
viscosity-shear rate behavior extending into the Newtonian range for
samples of such high viscosity is apparently quite unusual. Capillary
instruments simply cannot achieve the low shear rates required to reach
70 for high molecular weight samples at low temperatures. It should also
be noted that the onset of shear rate dependence is very well-defined in
narrow-distribution systems. Studies with broad-distribution polymers,
VISCOSITY-SHEAR RATE 1781

To C
Fig. 10. Ratio of Rouse to experimental relaxation time, reduced to 160"C, as a
function of temperature for all linear polystyrenes. Symbols: (0)S-109 Kepes; ( A )
8109 Weissenberg; ( 0 )PC-4; ( 0 )L-1; (V)L-2.

where the shear sensitivity is smaller, might find nonsuperposition some-


what more difficult to establish.
A shear stress dependence in E, has been reported for one other polymer
system, lowdensity p ~ l y e t h y l e n e . ~This
~ could be caused either by anom-
alous temperature variations of T O with no change in the master curve
form, as in this work, or by temperature-dependent changes in the form of
the master function itself. It is not clear which of these two possibilities
is responsible.
An important question is whether these observations imply a corre-
sponding violation of conventional temperature superposition in linear vis-
coelasticity. As pointed out earlier, the long relaxation times in the linear
spectrum control the onset of shear rate dependence in many polymers.
Markovitz30 has shown that, if all relaxation times have the same tempera-
ture coefficient, and 71 is a characteristic or mean relaxation time in the
linear spectrum, then the temperature dependences of 7 0 ~ ~and 1 , the steady-
state shear compliance JZ for any sample should be related, such that :
TI =CvoJ; (7)
in which C is a constant, independent of temperature. Assuming a direct
correspondence between T O and the linear viscoelastic relaxation times,
the conventional superposition principle (eq. (3)) requires th a t the tem-
perature dependence of JZ be of the form JX a (pT)-'. Thus JZ should
be quite insensitive to temperature, rising very slowly with decreasing tem-
perature. The observation here th at T O rises more slowly with decreasing
temperature than 70 implies (still assuming the correspondence between T O
1782 PENWELL, GRAESSLEY, AND KOVACS

and T I ) the reverse: that J: decreases with decreasing temperature below


190C.
~ reported that J: in polystyrene melts begins
Plazek and O R ~ u r k ehave
to fall as T o is approached from above. However, this effect seems to
appear only below about 140C, considerably below the range 150-190C
where T O behaves anomalously. Barlow and co-uorkers31have found what
seem t o be similar anomalies in the temperature coefficient of compliance
for small molecule liquids near T,. Again such results are obtained rather
closer t o T othan in the present work.
It seems unlikely that deviations of this magnitude from the conven-
tional superposition principle would have been overlooked in the extensive
linear viscoelastic studies which have been performed on pOlystyrene melts
over the temperature range 150-190C.14~28Both instruments used in this
study are plate-cone viscometers, so it is possible that some unsuspected
pecularity associated with this geometry and active only in highly viscous
systems, is causing the anomaly. This seems unlikely in view of the agree-
ment in Figure 1 among data obtained with different cone angles and platen
diameters, but we plan t o study the same systems in capillary flow to see
whether the anomaly persists.
For the time being we will set aside these concerns and conclude that the
temperature anomaly in T O is real and related to the same factors which
cause Plazeks anomaly in JE. It is assumed that these factors are some-
how exaggerated in the case of shear rate dependence in the viscosity and
therefore persist to somewhat higher temperatures. We believe i t is only
a coincidence that both the limit for the WLF form in 70 and the termination
of anomalous temperature dependence in T Ooccurs around 190C.
The fundamental explanation for the temperature anomaly, and whether
i t occurs in polymers other than polystyrene, is not known. We assume
that i t is somehoJv related t o the location of T , rrlative t o the temperature
of measurement. We do not know if it is characteristic of polymer melts
alone, or whether it can also appear in polymer solutions when measure-

B
B
OD

0.1 I I I I , / , , I I I I , , , ,

100 1000 10,000

Fig. 11. Reduced viscosity vs. reduced shear rate for a solut.ion of PC-4 in 72-butyl-
benzene. The concentration is 0.255 g/ml; the symbols are: I? 30C and 0 60C.
Data taken from Ref. 3.
VISCOSITY-SHEAR RATE 1783

ments are made correspondingly near the T, of the solution. It is certainly


not a n intrinsic property of the molecules, since Figure 11 shows almost
perfect superposition of v/qo versus toy/pT for a concentrated solution of
PC-4 in n-butylbenzene over the range 30-60C. However, T, for this
solution is probably somewhat below -100C. If either T - T, or TIT,
is the controlling variable, the solution at 30C is probably already well
into its limiting high-temperature behavior.
The help of Mr. Richard Crawley in the initial phases of the experimental work is
much appreciated. We are also grateful to the National Science Foundation (Grant
GK-34363) and the Northwestern University Materials Research Center for support of
this work.

References
1. G. C. Berry and T. G. Fox, Adv. Polymer Sci., 5 , 261 (1968).
2. R. A. Stratton, J . Colloid Interfacial Sci., 22, 517 (1966).
3. W. W. Graessley, R. L. Hazleton, and L. R. Lindeman, Trans. SOC.Rheol., 11,
267 (1967).
4. W. M. Prest, Jr., R. S. Porter, and J. M. OReilly, J. Appl. Polym. Sci., 14, 2697
(1970).
5. F. Bueche and S. W. Harding, J. Polym. Sci., 32, 177 (1958).
6. W. W. Graessley, J. Chem. Phys., 43,2696 (1965).
7. P. E. Rouse, J . Chem. Phys., 21, 1272 (1953).
8. J. D. Ferry, Viscoelustic Properties of Polymers, 2nd ed., Wiley, New York, 1970.
9. T . W. Dewitt, H. Markovitz, F. J. Padden, Jr., and L. J. Zapas, J . Colloid Sci.,
10, 174 (1955).
10. G. Kraus and J. Gruver, J. Polym. Sci., A-2,797 (1964).
11. W. W. Graessley and L. Segal, AIChE J., 16,261 (1970).
12. W. C. Uy and W. W. Graessley, Macromolecu~es,4,458 (1971).
13. M. Sakai, T. Fujimoto, and M. Nagasawa, Macromo/ecules, 5 , 786 (1972).
14. W. F. Knofi, I. L. Hopkins, and A. V. Tobolsky, Macromolecules, 4, 750 (1971).
15. D. J . Plazek and V. M. ORourke, J . Polym. Sci. A-2,9,209 (1971).
16. A. B. Bestul and H. V. Belcher, J . Appl. Phys., 24, 696 (1953).
17. R. L. Ballman and R. H. M. Simon, J . Polym. Sci., A-2, 3557 (1964).
18. G. Kraus and J. Gruver, J. Appl. Polym. Sci., 9,739 (1965).
19. R. A. Mendelson, Trans. SOC.Rheol., 9, 53 (1965).
20. R. S. Porter and J. F. Johnson, J . Polym. Sci. C, 15, 365 (1966).
21. R. A. Mendelson, Polym. Letters, 5 , 295 (1967).
22. A. Casale, R. S. Porter, and J. F. Johnson, J . Macromol. Sci.-Revs. Macromol.
Chem., C5, 387 (1971).
23. A. Kepes, J. Polym. Sci., 22,409 (1956).
24. K. H. Lee, L. G. Jones, K. Panddai, and R. S. Brodkey, Trans. Soc. Rheol., 14
555 (1970).
25. N. Nemoto, Polym. J . (Japan), 1, 485 (1970).
26. W. W. Graessley, J . Chem. Phys., 47, 1942 (1967).
27. R. C. Penwell and W. W. Graessley, J . Polym. Sci., Polym. Phys., Ed., 12, 213
(1974).
28. S. Onogi, T. Masuda, and K. Kitagawa, Macromo~ecu/es,3, 109 (1970).
29. R. S. Porter, J. R. Knox, and J. F. Johnson, Trans. SOC.Rheol., 12,409 (1968).
30. H. Markovitz, J. Phys. Chem., 6 9 , 671 (1965).
31. A. J. Barlow and A. Erginsav, PTOC. Royal SOC.(London), A327, 175 (1972).
Received April 22,1974

You might also like