Você está na página 1de 20

Transverse Cracking of M40J/PMR-II-50

Composites under ThermalMechanical


Loading: Part II Experiment and
Analytical Investigation

JAEHYUNG JU,* ROGER J. MORGAN AND TERRY S. CREASY


Department of Mechanical Engineering, Texas A&M University
College Station, TX 77843-3123, USA

E. EUGENE SHIN
NASA Glenn Research Center at Lewis Field, 21000 Brookpark Rd.
Cleveland, OH 44135, USA

ABSTRACT: In this study, the effects of thermal cycling combined with mechanical
loading on the microcracking of M40J/PMR-II-50 are investigated. Characterization
of the failure mechanisms are conducted based on the critical parameters which cause
composite microcracking, as presented in Part I. Based on the test results in Part I,
the tests with intermediate in-plane lamina strain (0.1750.350%) and an increased
number of thermal cycles are added. Elevated temperature thermal cycling
(23250 C) is also added to the original test plan to investigate the thermal cycling
temperature amplitude effect on microcracking of the composites. Observations
indicate that the elevated temperature exposure under mechanical loads causes an
easy fiber/matrix debonding. Subsequent exposure to cryogenic temperatures results
in fiber/matrix debonding due to the high thermal stresses associated with fiber/
matrix thermal expansion mismatch. Crack propagation under cryogenic exposures
is shown to be dominant with an increasing number of thermal cycles, especially
when combined with high temperature exposure associated with high amplitude of
cyclic thermal stresses.

KEY WORDS: M40J/PMR-II-50 (carbon fiber/polyimide composites), thermal


cycling, microcracks, interfacial failure, high temperature failure.

*Author to whom correspondence should be addressed. E-mail: jaehyung@tamu.edu


Figures 115 appear in color online: http://jcm.sagepub.com

Journal of COMPOSITE MATERIALS, Vol. 41, No. 9/2007 1067


0021-9983/07/09 106720 $10.00/0 DOI: 10.1177/0021998306067260
2007 SAGE Publications

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1068 J. JU ET AL.

INTRODUCTION

IGH TEMPERATURE THERMOSET composites have been recommended in the


H fabrication of high performance propulsion rocket engines and airframes in
supersonic transport aircrafts (SST) and missile airframes due to their excellent thermal
performance [1,2]. High performance propulsion rocket engines are exposed to
temperatures as high as 316 C (rapid heat-up rates up to 93 C/s).
The PMR (in situ polymerization of monomer reactants) polyimide resin has been a
candidate for these applications. PMR is an addition type polyimide, derived from
preformed oligomers, which undergoes thermal cross-linking or chain extension to form a
thermoset [3]. High temperature durability may allow PMR to be used at the anticipated
service temperatures of 300350 C. PMR-15 is the first generation PMR-type polyimide
developed by NASA and is recognized as a state of the art high-temperature matrix for
service temperatures of 300 C for 1000 h [4]. PMR-15 has excellent thermooxidative
stability and it retains its mechanical properties even at high temperatures. Second
generation PMR (PMR-II) was developed to meet the requirements for higher service
temperatures in advanced supersonic aircraft structures, tactical missile airframes, and
jet engines [5]. The PMR-II polyimide resins exhibit enhanced service temperature in the
range of 360380 C. The PMR-II-50 has been considered for high temperature and
high stiffness space propulsion composite applications such as the face-sheets of sandwich
structures with high stiffness carbon fibers [6].
Thermoset/carbon-fiber composites have been studied for use on propellant tanks of
reusable launch vehicles (RLVs) in an effort to reduce the gross liftoff weight [6,7]. They
are exposed to both cryogenic temperatures by the liquefied fuel and high temperatures by
the reentry heat flux. Simulated temperature amplitudes during flight have been conducted
with temperatures from 253 to 127 C with IM7/977-2 (carbon fiber/epoxy) composites
[8]. High temperature polymer composites are also candidate materials for a use in RLV
fuel tanks to reduce or eliminate insulation layers.
Expected loading conditions of RLV polymer composite cryogenic tanks are:
(i) in-plane tensile loading by fuel pressure, (ii) mechanical fatigue by fuel use and refill,
(iii) thermal stress by coefficient of thermal expansion (CTE) mismatch caused by flight,
remaining fuel, and thermal properties of insulators, (iv) thermal cycling, and (v) thermal
cycling combined with mechanical loading. Experimental and analytical investigations
of the damage, both in lamina and laminate level under tensile loading [912] and fatigue
[1315] were carried out and their failure was analyzed using mathematical models.
Thermal stress [1619] and its cycling effect [2030] on polymer composite damage also
have attracted many researchers. Some researchers have been interested in combined
thermal cycling and the mechanical loading effect on composite damage [30].
Unlike failure by high thermal residual stresses at cryogenic temperatures, possible
damage from high temperature exposure on high-temperature-cured composites has
received little attention. Ju and Morgan showed crack initiation of Bismaleimide (BMI)
carbon fiber composites at elevated temperatures (250 C) under mechanical loading [30].
High-temperature induced microcracking could be another damage mechanism.
The objective of this research is to investigate the effect of the combined thermal cycling
and mechanical loading on microcrack damage as a function of (i) thermal cycling
temperature amplitudes, (ii) number of thermal cycles, and (iii) mechanical in-plane strains
imposed on the laminates, which are vitally needed to ascertain design and materials
criteria for future aerospace applications.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1069

EXPERIMENTAL METHOD

A flow chart showing the experimental and analytical investigation plans is shown
in Figure 1. From the 2k experimental analysis in Part I, the number of cycles, thermal
cycling temperature amplitudes, and mechanical in-plane strains proved to be the most
important factors. For more specific characterization of microcracking, an intermediate
level of the important factors was added to the synergistic test at a constant heating rate
(4 C/min). Based on the results of Part I, (i) 23250 C thermal cycling temperature
amplitude, (ii) 0.263% average mechanical in-plane strain, and (iii) 2, 4, and 6 thermal
cycles were added. Four samples for each additional test were used for repetition of tests.
A failure analysis was performed using a strength criteria based on the thermal residual
stresses and in-plane mechanical stresses by bending.
The same lay-up of M40J/PMR-II-50 studied in Part I, which is [90/0]1s, was also used
in Part II. The resins modulus and strength, which are important properties for transverse
cracking analysis, were measured at three temperatures (Figures 2 and 3). Then these
properties of PMR-II-50 are assumed to decrease linearly below the resins Tg (390 C).
In comparison with 23 C, the tensile modulus, tensile strength, and shear strength decrease
about 28, 34, and 14, respectively, when PMR-II-50 is at 250 C.

Thermomechanical Loading Conditions

Bending was used as the applied loading in this study. However, bending does not
directly simulate the loading conditions of an RLV fuel tank. Instead, biaxial loading

Development of a conduction heating-based thermal cycling apparatus combined with mechanical loading

Materials study
2k Experimental design
(M40J/PMR-II-50)

Part I
Thermal cycling combined with mechanical loading
196~23C and 196~250C thermal cycling
0 and 0.488% in-plane mechanical strain
1 and 8 thermal cycles
1 and 4C/min heating rate Characterization of damage using statistical analysis

Thermal cycling combined with mechanical loading Part II


(4C/min heating rate)
23~250C thermal cycling
0.263% in-plane mechanical strain Analytical investigation: Nonisothermal thermoelasticity
2, 4, and 6 thermal cycles

Identify the damage mode of composites under thermal cycling combined with mechanical loading

Figure 1. Flow chart of plans to identify microcracking mechanisms induced by combined thermal cycling
with mechanical loading.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1070 J. JU ET AL.
5

4.5

3.5 Tensile modulus


Shear modulus
Modulus (GPa)

2.5

1.5

0.5

0
0 50 100 150 200 250 300 350 400
Temperature (C)

Figure 2. Modulus change of PMR-II-50 with varying temperatures.

70

60 64.02
Ultimate tensile strength
50 Ultimate shear strength
Strength (MPa)

40
32.64
35.76
30 28.12
27.34
20

10

0
0 50 100 150 200 250 300 350 400
Temperature (C)

Figure 3. Strength change of PMR-II-50 with varying temperature.

is generally considered to represent the loading conditions of an RLV fuel tank under fuel
pressure. Bending was used as the applied loading in this study because bending was
acceptable when considering the developed apparatus which combines the conduction
heating type of thermal cycling with mechanical loading. Even though bending does not
represent the loading condition of the cryogenic fuel tank, in-plane strain (or stress) can be
derived from bending and the derived in-plane strain (or stress) conditions can be applied

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1071

100.33 mm 100.33 mm

16.55 mm
27.20 mm

100.33 mm

y y
y
72.82 mm

91.61 mm

Block #1 Block #2 Block #3


(High strain application) (Low strain application) (No strain application)

Figure 4. Dimensions of aluminum blocks.

to assess the strain (or stress) conditions of the initial cracks with combined thermal
cycling. Three aluminum blocks were prepared for generating three different mechanical
in-plane strains (Figure 4). Two blocks are cylindrical and one is flat. The composite plate
samples were bent to conform to the outer surface of the two cylindrical blocks, which
are blocks #1 and #2, which induced two different in-plane strains. The composite plate
placed on the flat surface of block #3 does not supply in-plane mechanical strain.
The mechanical loading sequence of the bent composite plates is illustrated in Figure 5.
One side of a specimen is pin-clamped on the aluminum block and a force is carefully
applied to the other side. The other side edge is roller-clamped for a longitudinal
constraint-free condition.
In-plane stress at the laminate level without considering the fabric architecture is used
in this study. When the specimens are bent, the deflection curve w(x) of the specimen can
be derived from the geometry of the aluminum blocks by applying a large deflection
condition.
From Figure 5, the deflection geometry associated with x and w(x) coordinates
are derived,
x r sin 
1
w r cos   r  

where  is the maximum vertical deflection.


The first and second derivatives of w(x) with respect to x in terms of  are expressed
as Equations (2) and (3) after using the chain rule,
dw
 tan  2
dx
d2 w 1
2
 : 3
dx r cos3 
A deflection versus moment kinematic equation, which is the EulerBernoullis
exact differential equation for the elastic curve, is taken from a reference and is shown
in Equation (4) [31],
!
d2 w=dx2 Mxx
  4
2 3=2 EI
1 dw=dx

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1072 J. JU ET AL.

x
w

d
x

Figure 5. Large deflection bending associated with block geometry.

where w is deflection curve; Mxx, bending moments; E, modulus; and I, moment of inertia
( bh3/12, b is width and h thickness of the specimen).
Mxx is derived from the obtained deflection curves. For the laminated beam theory, it is
assumed that Myy Mxy 0. Substituting Equations (2) and (3) into Equation (4) results
in constant Mxx along the length direction; 0.1433 [Nm] for block #1 and 0.0759 [Nm]
for block #2. The moment induced bending strains were 0.65 to 0.65% for block #1
and 0.350.35% for block #2. It should be noted that the strains were calculated with
respect to laminate thickness.
From the moment obtained from Equation (4), the laminate in-plane stresses, which are
a function of a distance from the mid-surface, z, are calculated by Equation (5),

Mz  
xx x,z Q11 D11 Q12 D12 5
b
where

Dij Dij  Bik A1


kl Bij
Z h=2
   
Aij , Bij , Dij 1, z, z2 Qij dz
h=2
E1
Q11
1  12 21

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1073

E2
Q22
1  12 21
21 E1 12 E2
Q12
1  12 21 1  12 21
Q66 G12

Qij is plane stress-reduced stiffness; Aij, extensional stiffness; Bij, bending-extensional


coupling stiffness; Dij, bending stiffness; Ei, modulus; and vij, Poissons ratio.
Expanding Equation (5) to the stress components of layer levels results in
8 9
> @2 w0 >
8 9k 2 >
3k >  >
>
 11 Q 12 >
> @x2 > >
>
< xx >
= Q 0 >
< >
=
2
yy 6   7 @ w 0
z4 Q12 Q22 0 5 
>
: >
; >
> @x2 > >
xy 0 0 Q 66 >
> >
>
>
> @ 2
w >
>
: 2 0 ; 6
@x2
2 3k 2  38 9
Q 11 Q 12 0 D11 D12 D16 > < Mxx >
=
z6 7 6 7
4 Q 12 Q 22 0 5 4 D12 D22 D26 5 0 :
b >
: >
;
0 0 
Q66 D
16 D 
26 D
66 0

Assuming that the temperature varies linearly through the thickness direction of the
composite sample, the temperature difference can be written as

T T0 zT1 7

where T0 is the difference between the average temperature of the plate and the reference
temperature and T1 is the change in temperature per unit thickness of the plate. The
average temperatures of the sample are assumed to be the average values of the measured
top and bottom surface temperatures during the conduction heating. However, during
the cooling cycle from room temperatures to 196 C, the average sample temperature is
assumed to be the same as the top (or bottom) surface temperature, because the thermal
diffusivity of the composite sample is calculated to be 0.403 mm2/s from 1650 kg/m3
composite density, 812.5 J/kg C specific heat capacity, and 0.54 W/m C transverse thermal
conductivity, which means it takes about 2 min for the mid-surface temperature of the
composite plate to reach within 2 C of 196 C. Bechel et al. [24] measured the mid-surface
temperature when cooling a 1.64 mm thick polyimide composite sample (IM7/5250-4),
which has similar thermal properties to the samples used in this study, to 196 C by
embedding 0.102 mm thick k-type thermocouple. The center of the sample reached within
5 C of 196 C in less than 1.5 min in their experiment.
Due to the symmetry of the lamina scheme, thicknesses, and lamina properties about the
laminate, the coupling between bending and extension is eliminated and the governing
equation is simplified [32]. The symmetry reduced thermal force resultants are

8 T 9k 2 3k 8 9k
< Nxx = Z k1 Q 11 Q 12 0 < xx =
NT 4 Q 12 Q 22 0 5 yy T dz 8
: Tyy ; k  : ;
Nxy 0 0 Q66 2xy

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1074 J. JU ET AL.

where xx, yy, and xy are the coefficient of thermal expansion (CTE) in the xx, yy, and xy
directions, respectively and
8 T 9k 2 3k 8 9k
< Mxx = Z k1 Q11 Q12 0 < xx =
MT 4Q
12 Q22 0 5 yy Tz dz: 9
: Tyy ; k : ;
Mxy 0 0 Q66 2 xy

Once the thermal force and moment resultants are obtained, they can be converted to
a stress form for stress analysis. The force and moment resultants are just the thickness
averaged lamina forces and moments per unit width of the plate. Therefore, the thermal
T
axial stress in the kth lamina, f N gk is given by
8 T 9k 8 T 9
N
< xxT >
n T ok > = N =
1 < xx
N N
yy NTyy 10
>
: N ; > t : ;
xy
T
NTxy

where t is the laminate thickness.


T
The thermal bending stress in the kth lamina, f M gk , can be calculated by dividing the
force per unit width by the lamina thickness,
8 T 9k 8 T 9
M
< xxT >
n T ok > = M
1 < xx =
M yyM
k MTyy 11
:  MT >
> ; tD : T ;
xy
Mxy

where D(k) is the moment arm, D(k) z(k) (t/2).


Under the linear assumption, stress components of combined bending with thermal
loading are expressed as the sum of the mechanical and thermal load components

8 9k 8 NT 9k 8 MT 9k
>  > >  >
 Total k < xx >
> = < xxT >
> = < xxT >
> =
N M
 yy yy yy
>
: >
; >
> > > >
xy : NT >; : MT >
> ;
xy xy
2 3k 2  38 9
Q11 Q12 0 D11 D12 D16 >< Mxx >
=
z6 7 6 D D
4 Q12 Q22 0 5 4 12 22 D26 7
5 0 12
b >
: >
;
0 0 Q66 D16 D26

D66 0
8 9 8 9
>
> NTxx >
> >
> MTxx >
>
1 < = 1 < =
NTyy MTyy :
t>> > t  Dk > >
: NT > ; : MT >
> ;
xy xy

Mechanical and thermal residual in-plane stresses of blocks #1 and #2 through the
specimen thickness (0.5 mm) are shown in Figure 6. Total in-plane stresses by thermal
residual stresses and bending stresses are shown in Figure 7 corresponding to the two
different block radii at 196, 23, and 250 C. It should be noted that stress gradients exist

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1075
0.3 57.85 0.3 0.40
31.15 0.73
Block #2
Block #1 0.2 0.2 Block #2
16.20 30.08 Block #1

Thickness (mm)
Thickness (mm)

0.1 0.38 0.1 0.21


Stress (xx) (MPa) Stress (yy) (MPa)
0 0
800 600 400 200 0 200 400 600 800 100 80 60 40 20 0 20 40 60 80 100
0.1 0.1
0.2 0.2
0.3 0.3
(a) (b)
66.34 107.08
21.52 0.3 101.49 Tref = 371C 0.3
196C 34.73 196C
23C 0.2 163.80 0.2 23C
250C 250C
0.1 0.1
Thickness (mm)

Thickness (mm)
Thermal residual stress (xx) ( MPa)
0 0
800.00600.00400.00200.00 0.00 200.00 400.00 600.00 800.00 800 600 400 200 0 200 400 600 800
0.1 0.1 Thermal residual stress (yy) (MPa)
0.2 0.2

0.3 0.3

(c) (d)

Figure 6. In plane stresses by bending and thermal residual in-plane stresses: (a) in-plane stress (xx) by
bending; (b) in-plane stress (yy) by bending; (c) residual thermal in-plane stress (xx); and (d) residual thermal
in-plane stress (yy).

107.81 0.3
196C 79.370.3 124.19 159.34 164.53 35.46 196C
23C 0.2 0.2 23C
131.57 35.11
Thickness (mm)

250C 51.60 164.18 250C


0.1
Thickness (mm)

0.1 96.42 -107.46


Total stress (xx) (MPa) Total stress (yy) (MPa)
0 0
800 600 400 200 0 200 400 600 800 800 600 400 00 0 200 400 600 800
0.1 0.1
0.2 0.2
0.3 0.3

(a) (b)

52.67 97.49 107.47


0.3 132.64 164.20 0.3 35.12
196C 196C
23C 0.2 0.2 23C
250C 37.71 34.93 250C
117.68
Thickness (mm)
Thickness (mm)

0.1 164.01 0.1


82.54 Total stress (xx) (MPa) 107.28 Total stress (yy) (MPa)
0 0
800 600 400 200 0 200 400 600 800 800 600 400 200 0 200 400 600 800
0.1 0.1
0.2 0.2
0.3 0.3

(c) (d)

Figure 7. Total in-plane stresses for the two different bending conditions: (a) total stress (xx) for block #1;
(b) total stress (yy) for block #1; (c) total stress (xx) for block #2; and (d) total stress (yy) for block #2.

through the thickness of each layer due to the curvature associated with both distance
from mid-surface and in-plane ply modulus. It should also be noted that the thermal
residual stress at 23 C is higher than at 250 C, because the stress-free temperature is set to
be 371 C, which is the last cure temperature.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1076 J. JU ET AL.

250C 250C with


considering
3 transverse strength
2.6526 and modulus
2.5 reduction

2 23C 196C 1.7766


R value

1.5 1.3828

0.9788 0.9977
1 0.8124
0.6263 0.6257
0.5231 0.531 0.4447
0.5 0.3902

0
3C

C
3 C

C
C
3 C

0 C

C
C
C
C

50
50
50
50
50
196

196

196
at 2

at 2

at 2

at 2
at 2
at 2

at 2
at 2
at 2
at -

at -

at -
rain

rain

rain

rain
rain
rain

rain
rain
rain
rain

rain

rain
e st

e st

e st

e st
e st
e st

e st
e st
e st
e st

e st

e st
plan

plan

plan

plan
plan
plan

plan
plan
plan
plan

plan

plan
l in-

l in-

l in-

l in-
l in-
l in-

l in-
l in-
l in-
l in-

l in-

l in-
nica

nica

nica

nica
nica
nica

nica
nica
nica
nica

nica

nica
cha

cha

cha

cha
cha
cha

cha
cha
cha
cha

cha

cha
me

me

me

me
me
me

me
me
me
me

me

me
No

50%

50%

50%
50%
No

No
50%
50%
No

50%

50%
0.3

0.6

0.6
0.3
0.6
0.3
0.3

0.6
75~

25~

25~
75~
25~
75~
75~

25~
0.1

0.3

0.3
0.1
0.3
0.1
0.1

0.3

Figure 8. R-value of upper 90 ply based on the quadratic interaction failure criteria corresponding
mechanical strains, cryogenic and high temperature induced thermal stress, and mechanicalthermal
combined effects.

Total in-plane stresses in the lower 90 ply show both tensile and compression stress
states depending on temperatures (Figure 8). Therefore, only the upper 90 ply (outer
surface ply) was chosen of the four to measure the crack density, because the stress state of
this ply is always tensile.

RESULTS AND DISCUSSION

Failure Analysis using the Quadratic Interaction Criterion

A failure criterion is applied to investigate the initial damage condition. One of


the typical layer failure criteria is the quadratic interaction criterion. It is similar to the
TsaiWu failure criteria and used on each of the plies in the laminate to determine
the failure at a lamina level. The quadratic interaction criterion is given by

   
R2 Fxx x2 2Fxy x y Fyy y2 R Fx x Fy y  1 0 13

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1077

where

1
Fxx 14
XXc
1
Fyy 15
YYc
1 1
Fx  16
X Xc
1 1
Fy  17
Y Yc
1 p
Fxy  Fxx Fyy 18
2

X is the tensile strength in the fiber direction; Xc, compressive strength in the fiber
direction; Y, tensile strength transverse to the fibers; Yc, compressive strength transverse to
the fibers;  x, stress in the fiber direction; and  y, stress transverse to the fibers.
The physical meaning of the R is

ultimate
R 19
applied

and the positive R is taken from Equation (13). The quadratic interaction criterion
states that if R > 1, then the applied stress components are below the failure level. If
R<1, the lamina may fail. However, in this study, due to the lack of information on
transverse lamina failure strength of the composite, it cannot be said that the failure
will occur at R<1. It can only be said that the lower R implies a higher chance of crack
formation. By looking at the stress components on the upper 90 ply, which is of interest
in this study, R-values are derived for each loading condition. Figure 8 shows the
R-values expected at 23, 196, 250 C, and at 250 C considering transverse modulus
and strength reduction, under three different mechanical in-plane strains due to bending
and thermal residual stresses. Experimental results show that initial cracking occurs
under 0.3250.650% mechanical in-plane strain at 23 C. Therefore, the corresponding
R-value, 0.5231 is considered to be the critical value to decide whether crack
initiation occurs.

Cracking after Initial Thermal Exposure

Crack densities for three thermal cycling conditions are shown in Figures 911.
The crack density under the cryogenic thermal cycling (196 to 23 C) is the lowest among
the three thermal cycling conditions. Crack densities under elevated temperature thermal
cycling (23 to 250 C) are 1319% higher than the crack densities under cryogenic thermal
cycling. Cryogenic and elevated temperature combined thermal cycling (196 to 250 C)
has the highest crack density, which is 4044% higher than the crack density under
elevated temperature thermal cycling.
Cooling to 196 C is known to be an important factor that leads to large thermal
stresses, which include microcracking. Moreover, composites, post-cured at high

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1078 J. JU ET AL.
0.800

0.700 RT to 196C
No mechanical strain
0.175~0.350% transverse strain
0.600
Crack density (1/mm)
0.325~0.650% transverse strain
0.500

0.400

0.300

0.200

0.100

0.000
0 1 2 3 4 5 6 7 8
Number of thermocycles

Figure 9. Crack densities as a function of increasing number of cycles (23 to 196 C).

0.800

0.700 RT to 250C

0.600 No mechanical strain


Crack density (1/mm)

0.175~0.350% transverse strain


0.500 0.325~0.650% transverse strain

0.400

0.300

0.200

0.100

0.000
0 1 2 3 4 5 6 7 8
Number of thermocycles

Figure 10. Crack densities as a function of increasing number of cycles (23250 C).

temperatures such as 371 C, experience more severe thermal stress conditions in both
the matrix and fibermatrix interface when the composites are exposed to cryogenic
temperatures. Interfacial debonding seems to be a failure mode at cryogenic
temperature as can be seen in Figure 12(a), which gives rise to the conclusion that
strength of interface is lower than that of the matrix or that the local stress level of the
interface is higher than that of the matrix. However, an explicit explanation of the
interfacial debonding is difficult because the interface (M40J between PMR-II-50) strength
is unknown up to now.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1079

0.800 RT to 196C to 250C


No mechanical strain
0.700 0.175~0.350% transverse strain
0.325~0.650% transverse strain
0.600
Crack density (1/mm)

0.500

0.400

0.300

0.200

0.100

0.000
0 1 2 3 4 5 6 7 8
Number of thermocycles

Figure 11. Crack densities as a function of increasing number of cycles (23 to 196 C to 250 C).

Future sites of debonding


Interfacial debonding

Interfacial debonding

Initial microcrack precursor

(a) (b)

New debonding sites


Latent crack propagation

Interfacial microcrack propagation


Debonding width increase (Connection of debonding sites)

(c) (d)

Figure 12. Optical micrograph image of M40J/PMR-II-50 transverse cracking under one cycle of thermal
cycling: (a) 23 to 196 C (b) 23 to 250 C; (c) 23 to 196 to 250 C; and (d) 23 to 250 to 196 C.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1080 J. JU ET AL.

In addition to the cure induced extreme thermal residual stresses associated with
cryogenic temperatures, cool-down degradation might occur, and is mainly caused by the
fact that thermal stresses cannot relax sufficiently in highly cross-linked matrix systems.
The thermal stresses also preload the matrix at low temperature and lower the critical
fracture strain, resulting in a decrease of strength.
Elevated temperature thermal cycling itself does not appear to be a factor causing
microcracking by the thermomechanical model based on thermal residual stress
components, assuming that the stress-free temperature is 371 C, which is the final
postcure temperature. But, initiation of microcracking was observed at the fiber
matrix interface as shown in Figure 12(b). Temperature dependent strength and modulus
should be considered for a failure criterion at high temperatures. Normally, polymers
tested at high temperatures show a lower strength threshold for damage initiation.
The PMR-II-50 matrix at high temperatures shows strength and modulus reductions
as shown in Figures 2 and 3. Therefore, the matrix tensile strength and the matrix
dominated transverse strength should be changed to 42.28 and 10.57 MPa, respectively,
and the quadratic interaction failure criterion calculation is repeated with the revised
strength values. The repeated calculation of the quadratic interaction failure criterion,
however, still shows that there is no chance of lamina failure at high temperatures
based on the thermal residual stresses (Figure 8). The specimen weight loss of 0.01%
after one thermal cycle from 23 to 250 C does not seem to be enough to explain the
high temperature damage associated with the resins thermal degradation on high
temperature exposures.
There are several possible failure modes of composites at high temperatures.
A possible strength reduction at the interface (fiber/matrix interface zone) gives rise to
weakness in bonding, resulting in interface failure at high temperatures. VanLandingham
et al. showed an interface strength reduction at an elevated temperature (200 C)
using nano indentation and finite element analysis [33,34]. They also showed a strength
gradient over the interface with increasing temperature. An epoxy compatible sizing
between PMR-II-50 and M40J could be a reason for interfacial failure at high
temperatures. Allred et al. characterized the chemistry of sizing on M40J, manufactured
from Toray, and PMR-II-50, which are exactly the same materials used in our study [35].
They found that the Toray sizings routinely added to commercial carbon fibers are
not compatible with PMR-II-50 high-temperature polyimide matrix resins. They also
found that the Toray sizing is very nonuniform on the M40J fiber surface after
characterizing with scanning electron microscopy (SEM). High local thermal stresses
at the interface might happen due to the rough fiber surface caused by the sizing effect.
The sizing material itself could degrade at high temperatures, resulting in weak
interfacial adhesion.
The combined cryogenic/elevated temperature thermal cycling increases the crack
density with maintaining interfacial failure mode as shown in Figure 12(c). The crack
density increase cannot be explained in terms of thermoelasticity which is based on thermal
residual stresses.
Cryogenic temperature (196 C) exposures after elevated temperature (250 C) showed
nearly 26% increase in crack density compared with the high temperature (250 C)
exposures after 196 C (Figure 13). Microcracks initiated by interfacial debonding
at 250 C are considered to be connected together at 196 C under brittle matrix
conditions (Figure 12(d)).

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1081

Figure 13. Comparison of crack densities between RT to 196 to 250 C and RT to 250 to 196 C.

Damage with Increasing Number of Thermal Cycles

Intermediate number of cycles (2, 4, and 6) were added to the original tests in Part I
and the measured crack densities are shown in Figures 911. All three thermal cycling
conditions show an increase of crack densities with an increasing number of thermal
cycles. The elevated temperature thermal cycling (23250 C) shows a slow increase in
crack density with an increasing number of cycles (Figure 10) compared to the other
thermal cycling conditions. Cryogenic temperature thermal cycling induced a higher crack
density than elevated temperature thermal cycling after eight cycles (Figures 9 and 10).
The combined cryogenic/elevated temperature thermal cycling accelerated the crack
propagation with an increasing number of cycles, resulting in 3845% and 4854% higher
crack densities than under cryogenic and elevated temperature thermal cycling after eight
cycles, respectively.
High thermal residual stresses are known to be the main damage mechanism under
cryogenic temperature exposures. Microscopy shows that cracks propagated easily
through the thickness by linkage of debonding sites associated with brittle conditions
at low temperatures (Figure 14(a)).
Crack density increases with an increase in the number of elevated thermal cycles and
are shown in Figure 10. An increase after eight cycles from 0.096 to 0.139 mm1 in crack
densities due to cycling effect is shown. Initial debonding related cracks hardly propagate
along the transverse direction. This may be due to greater matrix ductility at high
temperatures. The connection of debonding sites after eight elevated temperature thermal
cycles can be seen in Figure 14(b), but it is not as active as cryogenic thermal cycling.
A 0.4% weight reduction observed after eight thermal cycles with 10 min of 250 C
exposure (Figure 15) does not seem to significantly affect the thermal degradation of
resin associated with the chain breaking. Instead, easy mobility of chain molecules at

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1082 J. JU ET AL.

Increase in crack
length and width

Connection of debonding sites

(a) (b)

Crack propagation

(c)

Figure 14. Optical micrograph image of M40J/PMR-II-50 transverse cracking under eight cycles of thermal
cycling: (a) 23 to 196 C; (b) 23 to 250 C; and (c) 196 to 250 C.

100.05

100

99.95

99.9

99.85
Weight (%)

99.8

99.75

99.7

99.65

99.6

99.55
0 1 2 3 4 5 6 7 8 9
Number of thermocycles

Figure 15. Weight loss under high-temperature thermocycling (23250 C) with increasing number of
thermocycles.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1083

high temperatures might cause a loss in bonding strength between the fiber and the matrix
resulting in debonding at fiber/matrix interfaces [3638].
The combined cycling effect increases damage with an increasing number of cycles.
Once cracks are initiated, high temperature-induced interfacial debonding and cryogenic
temperature-induced thermalresidual stresses accelerate the damage. The crack density
increases from 0.301 to 0.387 mm1 after eight cycles of combined thermal cycling as
shown in Figure 11. Cracks are fully developed in the thickness direction with an
increasing number of combined thermal cycles as shown in Figure 14(c).
Bechel et al. [39] tried to explain the phenomena. They showed that room temperature
mechanical fatigue testing of a carbon/epoxy composite produced interesting results.
When a peak cyclic constant tensile loading was applied while varying the minimum stress,
laminate failure occurred only after a few cycles. The analogy between mechanical fatigue
and thermal cycling seems probable but still cannot explain the damage from room
temperature to 250 C cycle profile.
There were some endeavors to investigate the onset of microcracking using the
fracture mechanics based crack propagation criterion such as the strain energy release rate
under mode I opening, GI. Henaff-Gardin and Lafarie-Fernot [40] calculated the variation
of the strain energy release rate GI as a function of crack density which was obtained from
experiments for different temperature amplitudes: (i) 200 to 20 C, (ii) 200 to 50 C,
(iii) 200 to 90 C, and (iv) 200 to 130 C. They found that 200 to 130 C thermal cycling
showed a higher crack density than other three thermal cycling conditions both in initial
cracking and propagation, which agrees with the present results under 196 to 23 C and
196 to 250 C.

CONCLUSIONS

Failure mechanisms were investigated under three thermal cycling conditions


(196 to 23 C, 23 to 250 C, and 196 to 250 C) with a developed conduction heating
type thermal cycling apparatus. Both cryogenic and high temperature exposures induced
initiation of fibermatrix interfacial microcracking; however, different failure modes
occurred.
High thermal residual stresses at cryogenic temperatures affected fibermatrix
expansion mismatch and high local thermal stresses, resulting in initiation of fiber
matrix debonding. Cryogenic thermal cycling appears to accelerate crack propagation
associated with the brittle matrix condition at low temperatures.
Epoxy sizing of M40J fibers could be affected by high temperature (250 C) thermal
degradation, resulting in interfacial adhesion failure. The high temperature thermal
exposure also appears to cause a drop in interface strength, resulting in interface failure
by the applied transverse mechanical in-plane strain. A smaller increase in crack density
with an increasing number of elevated temperature thermal cycles compared with
cryogenic thermal cycling can be explained by the higher ductility of PMR-II-50 matrix
at high temperatures.
Finally, initiation of cracking is more dominant at high temperatures than at low
temperatures. Subsequent exposure to high thermal residual stress at cryogenic
temperatures after high temperature exposures results in composite matrix cracking
associated with the fibermatrix thermal expansion mismatch and a brittle matrix
condition. The thermal cycling between 196 and 250 C shows acceleration of cracking

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1084 J. JU ET AL.

with an increasing number of thermal cycles related to higher cyclic thermal stresses
compared to cryogenic thermal cycling (196 to 23 C) and elevated temperature thermal
cycling (23 to 250 C).

ACKNOWLEDGMENTS

This research was kindly funded by Dr Charles Lee of the Air Force Scientific Office
of Research (AFSOR) and the Advanced Technology Program sponsored by the State
of Texas.

REFERENCES

1. Stokes, E.S., Shin, E.E. and Sutter, J.K. (2002). Mechanical Testing of PMCs under Simulated
Rapid Heat-up Propulsion Environments (I. Temperature Measurement), In: Proceedings of
47th International SAMPE Symposium, Long Beach, CA, May 1014, pp. 356370.
2. Shin, E.E., Sutter, J.K., Eakin, H., Inghram, L., McCorkle, L., Scheiman, D., Papadopoulos, D.
and Kerze, F. (2002). Design and Fabrication Issues of High Temperature PMCS for Aerospace
Propulsion Applications, In: Proceedings of 47th International SAMPE Symposium, Long Beach,
CA, May 1014, pp. 314355.
3. Scola, D.A. (2002). Polyimide Resins, In: ASM Handbook-composites, Vol. 11, pp. 78115.
4. Serafini, T.T., Delvigs, P. and Lightsey, G.R. (1972). Thermally Stable Polyimides from
Solutions of Monomeric Reactants, Journal of Applied Polymer Science, 16(4): 905915.
5. Poveromo, L.M. (1987). In: Serafini, T.T. (ed.), High Temperature Polymer Matrix Composites,
Noyes Data, Park Ridge, NJ, pp. 320.
6. Achary, D.C., Biggs, C.G., Bouvier, C.G., McBain, M.C. and Lee, W.Y. (2005). Composite
Development and Applications for Cryogenic Tankage, In: 46th AIAA/ASME/ASCE/AHS/
ASC Structures, Structural Dynamics & Materials Conference, Austin, Texas, April 1821.
7. Rivers, H.K. (1999). Cyclic Cryogenic Thermal-mechanical Testing of an X-33/RLV Liquid
Oxygen Tank Concept, NASA/TM-1999-209560, Langley Research Center, Hampton, Virginia.
8. Kessler, S., McManus, H. and Matuszeski, T. (2001). The Effects of Cryocycling on the
Mechanical Properties of IM7/977-2, In: Proceedings of the American Society for Composites,
Blacksburg, VA, September 1214, pp. 93111.
9. Renard, J., Favre, J.-P. and Jeggy, T. (1993). Influence of Transverse Cracking on Ply
Behavior: Introduction of a Characteristic Damage Variable, Composite Science and Technology,
46(1): 2937.
10. Caron, J.F. and Ehrlancher, A. (1997). Modeling the Kinetics of Transverse Cracking in
Composite Laminates, Composite Science Technology, 57(910): 12611270.
11. Gudmundson, P. and Alpman, J. (2000). Initiation and Growth Criteria for Transverse Matrix
Cracks in Composite Laminates, Composite Science and Technology, 60(12):185195.
12. Rebiere, J.L. and Gamby, D. (2004). A Criterion for Modeling Initiation and Propagation of
Matrix Cracking and Delamination in Cross-ply Laminates, Composite Science and Technology,
64(1314): 22392250.
13. Kobayashi, S. and Takeda, N. (2002). Experimental and Analytical Characterization of
Transverse Cracking Behavior in Carbon/Bismaleimide Cross-ply Laminates under Mechanical
Fatigue Loading, Composite Part B, 33(6): 471478.
14. Henaff-Gardin, C. and Lafarie-Frenot, M.C. (2002). The Use of a Characteristic Damage
Variable in the Study of Transverse Cracking Development under Fatigue in Cross-ply
Laminates, International Journal of Fatigue, 24(24): 389395.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
Transverse Cracking of M40J/PMR-II-50 Composites 1085

15. Kawai, M., Yajima, S., Hachinohe, A. and Takano, Y. (2001). Off-axis Fatigue Behavior of
Unidirectional Carbon Fiber-reinforced Composites at Room and High Temperatures, Journal
of Composite Materials, 35(7): 545576.
16. Park, C.H. and McManus, H.L. (1996). Thermally Induced Damage in Composite laminates:
Predictive Methodology and Experimental Investigation, Composite Science and Technology,
56(10): 12091219.
17. Pagano, N.J., Schoeppner, G.A., Kim, R. and Abrams, F.L. (1998). Steady-state Cracking and
Edge Effects in Thermo-mechanical Transverse Cracking of Cross Ply Laminates, Composite
Science and Technology, 58(11): 18811825.
18. Kim, R.Y., Crasto, A.S. and Schoeppner, G.A. (2000). Dimensional Stability of Composite in a
Space Thermal Environment, Composite Science and Technology, 60(1213): 26012608.
19. Whitley, K. and Gates, T.S. (2004). Thermal/Mechanical Response of a Polymer Matrix
Composite at Cryogenic Temperatures, AIAA Journal, 42: 19912001.
20. McManus, H.L., Bowels, D.E. and Tompkins, S.S. (1996). Prediction of Thermal Cycling
Induced Matrix Cracking, Journal of Reinforced Plastics & Composites, 15: 124140.
21. Ahlborn, K. (1991). Durability of Carbon Fiber Reinforced Plastics with Thermoplastic
Matrices under Cyclic Mechanical and Cyclic Thermal Loads at Cryogenic Temperatures,
Cryogenics, 31(4): 257260.
22. Timmerman, J.F., Tillman, M.S., Hayes, B.S. and Seferis, J.C. (2002). Matrix and Fiber
Influences on the Cryogenic Microcracking of Carbon Fiber/Epoxy Composites, Composites
Part A: Applied Science and Manufacturing, 33(3): 323329.
23. Timmerman, J.F., Hays, E.S. and Seferis, J.C. (2003). Cryogenic Microcracking of Carbon
Fiber/Epoxy Composites: Influences of Fiber-Matrix Adhesion, Journal of Composite Materials,
37(21): 19391950.
24. Bechel, V.T., Fredin, M.B., Donaldson, S.L., Kim, R.Y. and Camping, J.D. (2003). Effect of
Stacking Sequence on Micro-cracking in a Cryogenically Cycled Carbon/Bismaleimide
Composite, Composites Part A: Applied Science and Manufacturing, 34(7): 663672.
25. Bechel, V.T. and Kim, R.Y. (2004). Damage Trends in Cryogenically Cycled Carbon/Polymer
Composites, Composite Science and Technology, 64(12): 17731784.
26. Owens, G.A. and Schofield, S.E. (1988). Thermal Cycling and Mechanical Property Assessment
of Carbon Fiber Fabric Reinforced PMR-15 Polyimide Laminates, Composite Science and
Technology, 33(3): 177190.
27. Shimokawa, T., Katoh, H., Hamaguchi, Y., Sanfongi, S. and Mizuno, H. (2002). Effect of
Thermal Cycling on Microcracking and Strength Degradation of High-temperature Polymer
Composite Materials for Use in Next-generation SST Structures, Journal of Composite
Materials, 36(7): 885895.
28. Henaff-Gardin, C., Lefarie-Frenot, M.C. and Gamby, D. (1996). Doubly Periodic Matrix
Cracking in Composite Laminates Part 2: Thermal Biaxial Loading, Composite Structure,
36(12): 131140.
29. Kobayashi, S., Terada, K., Ogihara, S. and Takeda, N. (2001). Damage-mechanics Analysis of
Matrix Cracking in Cross-ply CFRP Laminates under Thermal Fatigue, Composite Science and
Technology, 61(12): 17351742.
30. Ju, J. and Morgan, R.J. (2004). Characterization of Microcrack Development in BMI-carbon
Fiber Composite under Stress and Thermal Cycling, Journal of Composite Materials, 38(22):
20072024.
31. Timoshenko, S.P. (1953). History of Strength of Materials, McGraw-Hill, New York.
32. Reddy, J.N. (2004). Mechanics of Laminated Composite Plate: Theory and Analysis, 2nd edn,
CRC Press, Boca Raton, FL.
33. VanLandingham, M.R., Dagastine, P.R., Eduljee, R.F., McCullough, R.L. and
Gillespie, J.W., Jr. (1999). Characterization of Nanoscale Property Variations in Polymer
Composite System: 1. Experimental Results, Composites Part A: Applied Science and
Manufacturing, 30(1): 7583.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012
1086 J. JU ET AL.

34. Bogetti, T.A., Wang, T., VanLandingham, M.R. and Gillespie, J.W., Jr. (1999).
Characterization of Nanoscale Property Variations in Polymer Composite Systems: 2.
Numerical Modeling, Composites Part A: Applied Science and Manufacturing, 30(1): 8594.
35. Allred, R.E., Wesson, S.P., Shin, E.E., Inghram, L., Mccorkle, L., Papadopoulos, D.,
Wheeler, D. and Sutter, J.K. (2003). The Influence of Sizings on the Durability of High-
temperature Polymer Composites, High Performance Polymers, 15: 395419.
36. Fox, B.L., Hodgkin, J.H., Jar, B., Lowe, A. and Morton, T.C. (1999). Isothermal Ageing of An
Advanced High Temperature Carbon Fiber Composite, In: ICCM12, 12th International
Conference on Composite Materials, Paris, France, July 59, pp. 94102.
37. Otieno-Alego, V., Creagh, D., Jar, B., Fox, B.L. and Lowe, A. (2001). Characterization of the
Moisture Adsorption and Thermal Ageing Behavior of Polymeric Composite Systems Using
Raman Spectroscopy, In: Mallinson, B. (ed.), Ageing Studies and Lifetime Extension of
Materials, pp. 113122, Kluwer/Plenum Press, Dordrecht//New York.
38. Bowels, K.J., Papadopoulos, D.S., Inghram, L.L., McCorkle, L.S. and Klan, O.V. (2001). Long
Time Durability of PMR-15 Matrix Polymer at 204, 260, 288, and 316 C, NASA/TM-2001-
210602, NASA Glenn Research Center, Cleveland, OH.
39. Bechel, V.T., Camping, J.D. and Kim, R.Y. (2005). Cryogenic/Elevated Temperature Cycling
Induced Leakage Paths in PMCs, Composites Part B: Engineering, 36(2): 175182.
40. Henaff-Gardin, C. and Lafarie-Frenot, M.C. (2002). Specificity of Matrix Cracking
Development in CFRP Laminates under Mechanical or Thermal Loadings, International
Journal of Fatigue, 24(24): 171177.

Downloaded from jcm.sagepub.com at UNIV NORTH TEXAS LIBRARY on February 22, 2012

Você também pode gostar