You are on page 1of 9

Methatms o/CooqJosite Materuds. Vol. 33, Nn. 6.

1997

T I M E AND T E M P E R A T U R E DEPENDENT D E F O R M A T I O N O F
P O L Y ( E T H E R E T H E R K E T O N E ) (PEEK)

Robert D. Maksimov and Josef Kubat

The stress-strain behavior in tension and the effect of temperature on the creep of poly(ether ether ketone)
(PEEK) have been studied. At room temperature, - 130" below the glass-transition temperature, the material
does not become brittle, and the specimens show necking in tension over a wide range of elongation rates.
The stress and strain at yield and the strain at break are almost linear functions of the logarithmic elongation
rate. The values of stress and strain at yieM increase slightly with increasing elongation rate, while the strain
at break decreases markedly. The short-term creep tests were conducted at temperatures extending from 20 to
200"C. The glass-transition temperature was found to be about 155"C. The creep of PEEK is greatest at
temperatures above 130"C. In the glass region the time dependence of the deformation is much weaker. It
has been found that the time-temperature relation for PEEK corresponds well with its thermorheological
simplicity in the temperature range investigated. The data on the temperature shift factor below and above
the glass-transition temperature may be fitted separately to the Arrhenius and WiUiams-Landel-Ferry
(WLF) equations, respectively. The long-term creep tests show that PEEK has excellent creep resistance at
room temperature. After 14-month tests at a stress level of 30 MPa the total strain exceeds the instantaneous
elastic strain only by a factor of 1.15.

INTRODUCTION

Poly(ether ether ketone) (PEEK) is known as a high-performance thermoplastic polymer with excellent physical and
mechanical properties. This material was created in 1977 and its industrial production started in the 80s. Since PEEK has
outstanding electroinsulation characteristics, it was used initially for high-quality insulation, mainly in the electrotechnical
industry. Later on PEEK was used as a thermoplastic matrix for composites containing particulate fillers or reinforcing
fibers, especially carbon fibers [1-7]. The practical applications of PEEK have thus been expanded significantly, not only in
the automotive, chemical, aerospace, and medical industries, but also in mechanical engineering and home appliance produc-
tion.
PEEK is characterized by high values of its key mechanical parameters (strength, rigidity, impact toughness) and also
by high thermal, chemical, radiation, and environmental resistance. Low moisture absorption is a specific feature of both the
pure PEEK resin [81 and PEEK-based composites [9]. In particular, the authors of [9] found that carbon fiber-reinforced
PEEK composites show excellent mechanical properties even in a hostile environment (80"C. 85% RH).
PEEK is a crystalline polymer with a relatively low rate of crystallization. The degree of its crystallinity depends
essentiaUy on the rate of cooling, changing from 50% at a cooling rate of 3*C/rain to 25% at 1000*C/min [5]. PEEK can be
processed through injection and compression molding or extrusion. Since the material has a high melt viscosity, the process-
ing temperatures are relatively high (360-390"C):
In general, high melt viscosity and processing temperature are shortcomings of heat-resistant thermoplastic matrices.
Related to this problem are recent studies on blends of thermotropic liquid crystalline polymers (LCP) with less expensive
thermoplastic engineering plastics such as polypropylene [10-13], poly(butylene terephthalate) [10, 14, 15], polycarbonate

Institute of Polymer Mechanics. Latvian Academy of Sciences. 23 Aizkraukles St.. LV-1006 Riga. Latvia. Depart-
ment of Polymeric Materials. Chalmers University of Technology. S-412 96 Gothenburg. Sweden. Published in Mekhanika
Kompozimykh Materialov. No. 6. pp. 734-746. November-December. 1997. Original article submitted November 24. 1997.

0 !91-5665/97/3306-0517$18.00 ~ 1998 Plenum Publishing Corporation 517


12(

lOq

8(

4t

0 4 8 12 16 20 24 28

Fig. 1. Stress-strain behavior of PEEK in tension at different elongation rates: 0.05 (1),
0.5 (2), 5 (3), and 50 mm/min (4). The curves are shifted horizontally for clarity. The shift
constants A for curves 1, 2, 3, and 4 are 0, 4, 8, and 12%, respectively.

[10. 16-20], poly(ethylene terephthalate) [21], polystyrene [21, 22], and polysulphone [23]. In many cases processing of these
materials is accompanied not only by formation of self-reinforcing LC fibrils but also by a substantial decrease in viscosity of
the blend melt, even at low contents of the LCP components. This facilitates manufacture of products from such blends. The
corresponding effect of an LCP component on the properties of PEEK has been investigated to a limited extent only [24].
When applying PEEK (either pure resin or fiber-reinforced) as a structural material, it is necessary to comply with
requirements related to the temperature-time dependence of the mechanical properties of this material, particularly its
deformation under long-term loading. The data available [25] on the short-time creep of PEEK within a narrow temperature
range (120-150"C) show that the material exhibits a marked dependence of the deformation on time and temperature, and this
should be taken into consideration in designing particular structures, especially structural elements with high dimensional
stability.
The objective of this paper is to present a study of the stress-strain behavior of PEEK over a wide range of elonga-
tion rates. In particular, the effect of temperature on creep over a wide temperature range and the time-temperature relation
have been investigated. In addition, data on the long-term (one year) creep behavior are shown.

Experimental

The PEEK used in this study was obtained from the Erta Plastics Company. Belgium. The dumbbell-shaped samples
(5.4 x 7.0 mm in cross section) used for mechanical testing were cut from plates with dimensions of 100 x 100 x 5.4 mm.
Since all the samples should have similar thermal prehistory, they were subjected to the following thermal treatment. First.
the samples were exposed to 200"C for 1 h. then cooled to 130"C at a rate of 1 K/min and held at this temperature for 3 h.
The samples were then cooled down to 20"C at a rate of 1 K/rain. As follows from [25]. such a treatment lessens the effect
of physical aging on the results when studying creep at temperatures below the glass transition temperature.
Stress-strain measurements (tensile tests) were carried out using a universal testing machine MTS (MTS Systems
Corp.. USA): The testing temperature was 21 +_ I*C. The distance between the grips was 50 mm for all samples. The
tensile deformation diagrams were measured for the following elongation rates: 0 . 0 5 . 0 . 5 . 5 . and 50 mm/min. Consequently.
the strain rate at the test start was 0.001. 0.01. 0.1. and 1 mm/mm rain. The deformations were measured with an MTS
extensometer (model 632.11 C-20) mounted on the narrow section of the specimens. The strain values higher than the strain

518
TABLE 1. Characteristics of the Mechanical Properties of PEEK in Tension
at Different Elongation Rates

Rate of elongation, v, mm/min 0.05 i 0.5 " I 50


Stress at yield, n~, MPa 101 107 112 117
Strain at yield, ~:,j, % 6.1 6.8 7.5 8.3
Strain at break. ~:t,, % 18.0 16.5 15.8 14.2

~:b,~
try, M P a
18
115 1

o 17 Sy,%
110 9

o 2 16
/
8

105 I 15 7

100 L = '
-1
:
0 1
I 14
2
]

L o g (v, m m / m i n )

Fig. 2. Stress at yield a.v (I), strain at yield ~v (2), and strain at break ~b
(3) vs. log elongation rate.

at yield were determined from the distance change between the grips. The yield stress is defined as the first point at which
the tangent of the s t r e s s - s t r a i n curve becomes zero. The tensile modulus of elasticity was determined from the slope of the
initial part of the s t r e s s - s t r a i n curve.
The effect of temperature on the short-term creep was investigated at 17 temperatures in the range from 20 to 200~
using a test apparatus equipped with a thermochamber. The loading time was around 1 see. The creep curves were measured
during 1 h. Afterwards, the samples were completely unloaded and the creep recovery curves measured. It should be noted
that 24 h after unloading the residual strains were negligibly small, implying that the material exhibited inverse viscoelastic
deformation (time-recoverable creep) under the conditions used. Since the experiments were carried out in the region of linear
viscoelasticity, it was necessary to reduce the stress level when the temperature was increased. If at 20~ the stress value was
12.5 MPa, at 200"C it should be only 1.6 MPa. The total strains at the moment of specimen unloading did not exceed 0.6%
in any case.
The long-term creep tests were carried out in a room adapted specifically for this purpose at a temperature of 200C
maintained constant within 1 ~ The creep deformation was measured for ten thousand hours ( m o r e than a year). The tests
are still in progress.
S t r e s s - S t r a i n Behavior. The stress-strain diagrams for four values of the elongation rate are shown in Fig. 1.
Each curve is the average of data obtained using three specimens. For clarity the curves have b e e n shifted horizontally: the
amounts of the shift &e = A are indicated in the legend. The stress values were calculated as the ratio between the measured
tensile force and the initial cross section of the specimens.
We note that three characteristic regions can be distinguished on the s t r e s s - s t r a i n curves. In the first region the
shape of a specimen does not vary and the deformation is uniform. When the strain reaches the yield point (Oy, gy), a local
constriction (necking) with a subsequent failure of the load appears. The constricted region starts to grow in both directions
towards the specimen clamps. However. this process does not take place along the whole narrow section of the specimen.
The crack occurs in the drawn part. and as a result there is a decrease in the load with a subsequent rupture of the specimen.

519
4 - E, G P a
~ ~ "o ..o..ql. ~
9 .o Q ,o -o 9

QO..Q...Q

! I I
T, ~
I I

0 40 80 120 "160 20O

Fig. 3. Quasi-static elastic modulus of PEEK as a function of temperature.

Some characteristics of the material behavior in tension at four different elongation rates are listed in Table 1. Figure
2 shows the stress and strain at yield and the strain at break as functions of the logarithmic elongation rate. As can be seen,
the characteristics presented are almost linear functions of log v. The values of try and ey increase slightly with increasing
elongation rate. For v = 50 mm/min these values are, respectively, 1.16 and 1.36 times greater than for v = 0.05 mm/min.
In contrast, the strain at break decreases with increasing v; for v = 50 mm/min it is 1.27 times less than for v = 0.05
mm/min.
It is important to note that these tests were done at a temperature of 21"C, i.e., about 130"C below the glass-tran-
sition temperature. At such a low temperature the material does not become brittle for a wide range o f elongation rates. This
is particularly important in engineering applications of the material.
Effect of Temperature on Short-Term Creep. The temperature dependence of the quasi-static elastic modulus was
determined prior to creep testing. The modulus was obtained as the ratio of stress to strain under rapid quasi-static loading
and unloading, the duration of which was about 1 see. This step-by-step procedure was repeated periodically at temperatures
increasing slowly from 20 to 200"C at a rate of 1 K/min. The data obtained are presented in Fig. 3. The tests show that the
elasticity modulus E decreases slightly in the range from 20 to 140~ However, a dramatic decrease of E is observed in the
range from 140 to 170*C. It follows from Fig. 3 that the value of the glass transition temperature Tg defined by the tempera-
ture at which the derivative dE/dT has its maximum value is about 155"C.
Figure 4 presents the averaged experimental creep compliance data for 17 temperature levels in the range from 20 to
200"C in a double logarithmic plot. The creep deformation was measured for 1 h. The creep data are represented as creep
compliance D(O = e(t)/a, where D(t) is the compliance, t the time, e and a the strain and stress, respectively.
As can be seen in Fig. 4. the inelastic strain of PEEK becomes substantial at temperatures above 130~ The creep
compliance after deformation for one hour at 200"C is higher than the compliance at 20~ by a factor of 8.9. The time
dependence of the deformations is much weaker in the glass region. However, even in this region one should not neglect
viscoelastic behavior of the material. With increasing time and temperature, the creep effect in the glass region naturally
increases too, even at temperatures much lower than Tg.
When testing the validity of time-temperature superposition for the material examined, it is useful to consider the
data shown in Fig. 4. The slope of the compliance curves suggests that a master curve may be constructed by shifting the
curves measured at different temperatures along the log time axis until they all superimpose on the curve for the reference
temperature chosen. Figure 5 presents such a master curve for the reference temperature 155~ As can be seen, the
procedure of time-temperature superposition to obtain a master curve can be performed without difficulty. This implies that
the compliance curves at different temperatures differ only by the time scale, that is,

D(iog t, T) = D(log t - log aT) = D[Iog (t/ar) ].

where a T is the time-temperature shift factor.

520
Log (D, MPa "1)
-2.6 .... ~...v.-~ .... ~ , . - v . ~ - v 200
.... v ...... ~" . .&,....A... ak - ' ' ' ' & "
-- 9 ~ 9 190
I ..V''""
........., 9 ..... . ,a...,,.,a 9 1 8 0
.A ." ""''&'" "'" ..,.~ . ~ -
. .....A."
.....

..a .... ='" 170


..111"" n 9 "it"i"
A-'"
...if
-2.8 ~ l ~ i ~ ~
9" .~..o 9 o 165
... 9 .11"" ..ID 9 9149

........ - "~' o.e 160


99 . 0"" .ll 9
..11. o"
..13"" .-'l'"
.'" 9
9 .-" 9
9 9 ..
I1"
..{3" ...0'
.-
-3.0 9 n "9
..- e"" ..0,0 "0 155
.-
9 0" ..0"
~
~]' 9 9" " .0""
.. 9 ..0"

...IS 9 ..'0""

.0."
9 ..- 9149 1 5 0
.-'0" .0

.9 9O " ...4r
Q." .9 .~-"
D
-3,2 .0 o"

9,'" 145
..-0" 9 . 0 . 0 "0
....-~" ....0'
..
0"" ..0"
.9 9.o v.-v 1 4 0
.. 9149 "9 9" ~ ' " ..v"
..0."" 9. 0 " " .... 9
,... 9 ..O'"" " ..V .''1r
.9 9 ... ....T. 9
-3.4 -e'" ..-o ........ .v ..... ~-~rv 1 3 0
....... o ...... . .......... 9 ..... v....~..,~....-v'" "120
O" ...... _ ......... 9 -- -.'V ......... -- .A. A 9 at" ~ '
-" ....
..... - .V .......
..... ~" - ..... Jr' 9149
9 ,.---" ~, ,~A
100
# . ....... .. ......... .. .... : ~"2......:::.e=9 . . . ... .. ......... .. ...... 9~ ......
"'"'""~
-e
..... " ....
...... ~~ "
u ''n 9149
80
/-~. . . . . . . . . . . . . . ...... ii ........ 9
:B ............... 9 ......... 9 .........
- D 9 .o....o.-t~-oa 60
9 . ...............
............. o9 ..........
........ 9o ...............
............... ~ ......
_c2. . . . . ~_
_ ........
. . . . . . . . ~....o. 9149 _ _
40
.............. o ......... o ............... o ...... o ........ o . . - . o - - o - . 9 -ooo 20

-3.6 I I I I I, I
1 2 3 Log (t, s)
Fig. 4. Experimental tensile creep compliance curves in logarithmic coordinates at temperatures f r o m 20
to 2 0 0 ~

The experimental values of the shift factor log a T are shown by dots in Fig. 6 as a function of temperature. Here the
curve of the temperature dependence of log a T shows several noticeable changes in the slope. This is due to inclusion of a
rather wide temperature range including the glass region (from 20 to 155~ and the region o f glass-to-rubber transition
(above 155*C).
Let us first consider the data obtained at temperatures above Tg. In numerous papers it has been reported that the
temperature shift factor in the temperature range from Tg to approximately (Tg + 100*C) for many types of polymeric
systems obeys a relation known as the Williarns-Landel-Ferry (WLF) equation [26]

log aT = { - q ( r - T~)}/(c2 + T - Tg),

w h e r e c I and c 2 are constants.


Using the method o f least squares for the data above the glass-transition temperature ( 1 5 5 < T < 2 0 0 ~ w e found
that the values o f c t and c 2 are 16.6 and 75.4. respectively. The calculated curve is s h o w n in Fig. 6 as a solid line. The
accuracy o f the approximation used may be considered quite satisfactory with a relative m e a n - s q u a r e error o f 1 9149

521
-26 L Log (D. MPa -1)

- 2 8 --

,t
-30 --
cp

-3.2
/

-3.4
/
-3.6 I Log (t I aT) i
-10 -5 0 5 10

Fig. 5. Creep compliance master curve for PEEK at the reference temperature 1550C. The symbols are
the same as in Fig. 4.

16 - Log a T
".
12 =.

8 '..... !l I
4 ""=""" . I .

..4
II i x-'-
-8 , , , T011 "3T, ~
40 80 120 160 200
Fig 6. The time-temperature shift factor log a T as a func-
tion of temperature. Dots are the experimental results. The
lines were calculated using the WLF equation with coeffi-
cients calculated in the present work (1) and universal
constants (2).

It should be noted that the calculated values of the coefficients c I and c 2 differ somewhat from those called universal
constants of the WLF equation, namely, c I = 17.44 and c 2 = 51.6 [26]. The curve calculated using the universal constants
is shown in Fig. 6 as a d a s h - d o t line. It is evident that the WLF equation with universal constants failed to agree with
experimental data. It is also important to note that the use of the WLF equation with universal constants as well as with the
coefficients obtained above for extrapolation to the glassy region is not possible, the discrepancy between the experimental
and computed results becoming substantial (see Fig. 6). This agrees completely with the conclusion made in [26] that the
WFL equation cannot be used at temperatures below Tg, even if all relaxation times within the glassy region have approxi-
mately the same temperature dependence. In this case we may apply the time-temperature superposition method, though the
dependence of log aT(T) on T usually better corresponds to the modified Arrhenius equation [26]

522
16 Log a T

~
94

-8 L~
2.0
2 I
I
F ,
103/To 2.5 3.0
, 103/T, K
3.5
I

Fig. 7. The time-temperature shift factor log a T versus


reciprocal temperature. Dots are the experimental results.
The lines were calculated using the Arrhenius (1) and WFL
(2) equations.

log ar - ~--.~R - ,

where U is the activation energy, R the universal gas constant, T the absolute temperature, and TO a constant.
We have checked this assumption using experimental data for PEEK in the glassy region, that is, at temperatures
from 20 to 155"C. The result obtained is shown in Fig. 7, where log a r is presented as a function of the reciprocal absolute
temperature. The separate fitting of the Arrhenius equation to the data from the glassy region is shown by a straight line
marked 1. The activation energy is 57.6 kcal/mole, and the relative mean-square error of the approximation used is 13%.
Figure 7 shows also line 2 computed using the WFL equation with coefficients c I and c 2 calculated above.
Since the temperature dependence of the time-temperature shift factor for PEEK is not monotonic in a wide
temperature range comprising the glass region and the glass-to-rubber transition, it seems difficult to represent this depen-
dence analytically by means of one equation. The analysis shows there is reason to believe that these difficulties may be
alleviated by separate fittings of the modified Arrhenius equation to the data from the glassy region and the W L F equation to
the data from the glass-to-rubber transition region.
Long-Term Creep Behavior. As noted above, the long-term creep tests were carried out for 10,000 h (about 14
months) at a temperature of 20 + I*C. Figure 8 shows experimental data (dots) of the creep compliance D(t) = e(t)/o versus
log time. The curve represents the average of data obtained using three specimens. The stress was 30 MPa for all specimens
examined. Standard deviations were less than 8%:
The tests show that at the stress used the deformations due to creep are relatively limited. On the whole, the material
has excellent creep resistance at room temperature. After 14 months, the total strain exceeded the instantaneous elastic strain
by a factor of 1.15 only: the total strain at a stress of 30 MPa is 0.93% and the instantaneous strain 0.81%. After a 1.4-month
test the inelastic (creep) strain is thus only 0.12%. The tests are still in progress.

CONCLUSION

The stress-strain behavior of PEEK at room temperature is influenced by the elongation rates to a limited extent
only. The material does not become brittle even when the elongation rate is increased by three orders of magnitude. The
specimens show necking in tension. The stress and strain at yield and the strain at break are almost linear functions of the
Iogaritlmlic elongation rate. The values of stress and strain at yield increase slightly with increasing elongation rate. while the
strain at break decreases markedly.

523
3.1
DxlO4, MPa-1 /

,. ~ . , = . e " ~
~e"
2.9
e~'e~ ~ eel"
O.. ,I "Q

2.7

1 YR.
/
2.5 I I f I I | I ~ I
1 3 5 7 Log (t, s)

Fig. 8. Long-term (14 month) creep compliance of PEEK at 20~ as


a function of logarithmic time.

Short-term creep tests were conducted at temperatures from 20 to 200"C. The glass-transition temperature was found
to be about 155"C. Consequently, the tests were carried out in the glassy region as well as in the region of the glass-to-
rubber transition. As expected, the effect of time on the corresponding mechanical parameters in the glassy region was
limited only, though not negligible. The creep manifests itself strongly at temperatures above 130~ The creep compliance
after deformation during one hour at 200"C is higher than the compliance at 20"C by a factor of 8.9.
The time-temperature behavior corresponds well to thermorheological simplicity in the temperature range investi-
gated. The temperature dependence of the shift factor within a wide temperature range comprising the glassy region and the
glass-to-rubber transition is not monotonic and it appears that the analytical representation of this dependence should be made
separately for the data below and above Tg. It is demonstrated that the temperature shift factor below and above the glass-
transition temperature may be fired separately to the Arrhenius and WLF equations.
The long-term creep tests show that PEEK has excellent creep resistance at room temperature. After 14 months at a
stress of 30 MPa, the total strain exceeds the instantaneous elastic strain only by a factor of t. 15. The tests are still in
progress.

Acknowledgements. The research was supported by the Royal Swedish Academy of Sciences, Stockholm, and the
Latvian Council of Science, Riga.

REFERENCES

I. A. C. Duthie and J. Damon, "The design and manufacture of a graphite fiber-reinforced poly(ether ether ketone)
(PEEK) tetrahedron truss array," Polymer Eng. Sci., 31, No. 1, 1-5 (1991).
2. A. Lustiger, F. S. Uralil, and G. M. Newaz. "Processing and structural optimization of PEEK composites," Polymer
Composites, U . No. 1, 65-75 (1990).
3. K. B. Kwarteng and C. Stark, "Carbon fiber-reinforced PEEK (APC-2/AS-4) composites for orthopaedic implants."
SAMPE Quarterly, 22. No. 1. 10-14 (1990).
4. J. A. Barnes and F. N. C o g ~ e l l , "Thermoplastics for space," SAMPE Quarterly, 21, No. 2, 22-27 (1989).
5. G. Jeronimidis and A. T. Parkyn, "Residual stresses in carbon fiber-thermoplastic matrix laminates," J. Composite
Materials. 22, 401-415 (1988).
6. M. J. Bozarth, J. W. Gillespie, and R. L. MccuUough. "Fiber orientation and its effect upon thermoelastic properties
of short carbon fiber-reinforced poly(ether ether ketone) (PEEK)," Polymer Composites, 8, No. 2, 74-81 (1987).
7. D. J. Blundell and B. N. Osborn. "Crystalline morphology of the matrix of PEEK-carbon fiber aromatic polymer
composites. Part 2: Crystallization behavior." SAMPE Quarterly, 17, No. 1. 1-6 (1985).
8. G. Mensitieri, M. A. Del Nobile. A. Apicella. and L. Nicolais, "Time and temperature dependent sorption in poly-
ether-ether-ketone (PEEK)." Polymer Eng. Sci.. 29. No. 24. 1786-1795 (1989).

524
L). M. M. Chen-Chi and Yur Shih-Wen. "Environmental effects on the water absorption and mechanical properties of
carbon fiber-reinforced PPS and PEEK composites. Part II," Polymer Eng. Sci.. 31, No. 1, 34-39 (1991).
10. W. Brostow. T. Sterzynski. and S. Triouleyre, "Rheological properties and morphology of binary blends of a
longitudinal polymer liquid crystal with engineering polymers," Polymer, 37. No. 9, 1561-1574 (1996).
I1. M. T. Heino and J. V. Sepp~il~i. "Studies on compatibilization on blends of polypropylene and a thermotropic liquid
crystalline polymer," J. Appl. Polymer Sci., 48, 1677-1687 (1993).
12. A. Datta, H. H. Chen, and D. G. Baird, "The effect of compatibilization on blends of polypropylene with a liquid-
crystalline polymer," Polymer, 34, No. 4, 759-766 (1993).
13. R. D. Maksimov and T. Sterzynski, "Mechanical properties of blends of liquid crystalline copolyesters with
polypropylene," Mekh. Kompozim. Mater., 30, No. 4, 442-450 (1994).
14. K. Engberg, M. Ekblad, P.-E. Werner, and U. W. Gedde, "Thermal and mechanical properties of injection molded
blends of a liquid crystalline polymer and poly(butylene terephthalate)," Polymer Eng. Sci., 34, No. 17, 1346-1353
(1994).
15. R. D. Maksimov, T. Sterzynski, and L Garbarczyk, "Structure and properties of injection molded blends of liquid
crystal polymer (40 PET/60 PHB) with poiy(butylene terephthalate)," Mekh. Kompozitn. Mater., 32, No. 5,676-689
(1996).
16. M. R. Nobile, E. Amendola, L. Nicolais, D. Acierno, and C. Carfagna, "Physical properties of blends of
polycarbonate and a liquid crystalline copolyester," Polymer Eng. Sci., 29, No. 4, 244-257 (1989).
17. D. Berry, S. Kenig, and A. Siegmann, "The mechanism of skin-core morphology formation in e• of
polycarbonate/liquid crystalline polymer blends," Polymer Eng. Sci., 33, No. 23, 1548-1558 (1993).
18. Q. Lin and A. F. Yee, "Mechanical properties of in situ composites based on polycarbonate and a liquid crystalline
polymer," Polymer, 35, No. 16, 3463-3469 (1994).
19. K. Engberg, O. Stromberg, J. Martinsson, and U. W. Gedde, "Thermal and mechanical properties of injection
molded liquid crystalline polymer/amorphous polymer blends," Polymer Eng. Sci., 34, No. 17, 1336-1345 0994).
20. W. Brostow, M. Hess, B. L. Lopez, and T. Sterzynski, "Blends of a longitudinal polymer liquid crystal with
polycarbonate: relation of the phase diagram to mechanical properties," Polymer, 3, 1551-1560 (1996).
21. P. Zhuang, T. Kyn, and J. L. White, "Characteristics of hydroxybenzoic acid-ethylene terephthalate copolymers
and their blends with polystyrene, polycarbonate and polyethylene terephthalate," Polymer Eng. Sci., 28, No. 17,
1095-1106 (1988).
22. N. Ogata, T. Tanaka, T. Ogihara, K. Yoshida, Y. Kondou, K. Hayashi, and N. Yashicla, "Effects of the addition of
a liquid crystalline copolyester to polystyrenes on blending torque and mechanical properties of blends," J. Appl.
Polymer Sci., 48, 383-391 (1993).
23. V. G. Kulichikhin and N. A. Plat~, "Composite blends based on liquid-crystal thermoplastics," Vysokomol. Soedin.,
Ser. A, 33, No. 1, 3-38 (1991).
24. R. E. S. Bretas and D. G. Baird, "Miscibility and mechanical properties of poly(ether imide)/poly(ether ether
ketone)/liquid crystalline polymer ternary blends," Polymer, 33, No. 24, 5233-5244 (1992).
25. A. A. Ogale, "Creep behavior of thermoplastic composites," in: Thermoplastic Composite Materials, L. A. Carlsson
(ed.), Amsterdam-Oxford-New York-Tokyo, Elsevier (1991), pp. 205-232.
26. }. Ferry, Viscoelastic Properties of Polymers, Wiley. New York (1961).

525

You might also like