Sunteți pe pagina 1din 9

The thermal and mechanical properties of carbon, glass and Kevlar

fibre reinforced epoxy composites are discussed, with particular


reference to the behaviour of these materials at cryogenic temperatures.
The effects of production techniques and various fibre arrangements are
determined.

Fibre-epoxy composites at low temperatures*


G. Hartwig and S. Knaak
Keywords: low temperature techniques, composite materials, fibre-epoxy composites

Nomenclature

auc

Ultimate compressive stress

cr/crUT Fatigue endurance limit (tensile threshold)


E

Young's modulus

V]LS

Interlaminar shear strength

Fibre volume fraction

to

Fibre angle

fi,L/L

Integral thermal expansion

RT

Room temperature

Indices

Relative thickness of angle-ply

Matrix

Coefficient of thermal expansion

Fibre

e
eUT

Strain
Ultimate tensile strain

Composite

X
/x

Thermal conductivity
Poisson's ratio

Polymers

cr

Stress

PC

Polycarbonate

cruT

Ultimate tensile stress

PSU

Polysulfone

Fibre composites are attractive alternatives to


metals because of their high specific strength or
stiffness or their excellent fatigue behaviour. They are a
necessary supplement to metals because of their low
electrical and thermal conductivities, the latter being
related to strength or stiffness. Their disadvantage
arises from the weak polymeric matrix and results in
low interlaminar shear strength and low transverse
strength. At low temperatures some properties of the
heterogeneous fibre-matrix system are superimposed by
such peculiarities as additional thermal resistance at
boundaries (Kapitza effect) by which the thermal
conductivity is reduced. In the course of cooling the
different thermal contractions of fibre and matrix give
rise to thermal residual stresses and strains which
influence most of the mechanical properties. At low
temperatures the majority of currently used matrices
are brittle and do not allow relaxation of residual
stresses or stress concentrations to take place. In future
developments the application of special, low tempera-

ture ductile thermoplastics will improve the mechanical


properties.
Cryogenic applications of polymeric fibre
composites are mainly in superconductivity, space
technology and handling of liquefied gases. They
include superconducting generators or pulsed magnets
for fusion reactors which call for materials resistant to
fatigue and electrically resistive to eddy currents.
Support elements for liquefied gas containers have to
be optimized with regard to low thermal conductivity
and mechanical strength. Transport vessels and
components used in space technology should be light
weight and resistant to fatigue. In several applications
low thermal contraction is necessary. The cryogenic,
mechanical and thermal properties of important
polymeric fibre composites will be presented.
The dependence of mechanical properties on production
techniques was studied for carbon fibre composites.

*Dedicated to Professor Dr. W. Heinz on the occasion of


his 60th birthday,

The fibres used in this work were E-glass fibres,


high modulus carbon fibres (M40, Toray), high tensile

0011-2275/84/011639-09
CRYOGENICS. NOVEMBER 1984

Materials

$03.00 1984 Butterworth ~t Co (Publishers) Ltd.


639

Table 1.

Fibre properties

Fibres

T 300
Carbon fibre M 4 0 A
A S-4
Fibre glass (E-glass)
Kevlar 49

oUT II ,

8UTII

GPa

GPa

3.5
2.4
2.8
3.0-3.4
2.8

1.5
0.6

230
400
210
70
140

3.1
2.1

EFII

EF *,

GPa
~ 24
70
~ 11

* - a t T = 4.2 K
carbon fibres (T300. Toray and AS4, Hercules) and
Kevlar-49 fibres (Du Pont). The principal fibre
properties are shown in Table 1.
Material samples were produced by filament
winding, for tubes, as well as wet lay-up and prepreg
techniques. Three epoxy resin systems were used to
form the polymeric matrices. All three are based on
bi,phenol-A and manufactured by Ciba Geigy. The
system Cy 221/Hy 979 is a flexible laminating resin:
the two others, Ly 556/Ht 972 and My 740/Hy 917, are
rigid at room temperature and are used in prepreg
techniques.
Epoxy resins are cross-linked polymers and show
brittle behaviour at low temperatures. The tensile
fracture strain for flexibilized laminating resins is of
the order of BUTM ~ 2% at 4.2 K. The systems used in
prepreg techniques are much more brittle and exhibit
an even lower fracture strain. Some types of
thermoplastics (eg PC, PSU) are ductile in their
behaviour even at low temperatures and show a

Table 2.

fracture strain of BUTM ~ 4-6% at 4.2 IC They are


candidates for matrices of future composites for
cryogenic applications?
For all matrices not only their cryogenic
properties are decisive but at RT additional
requirements must be fulfilled. This applies mainly to
the creep characteristic of composites which should be
sufficiently low. However, in most cryogenic magnet
applications high loads are applied only at low
temperatures, where creep processes are negligible. At
RT they have to withstand the intrinsic weight load
only (eg structural parts for superconducting magnets
and generators).
For one- and two-dimensional load conditions
the following fibre arrangements were chosen:
unidirectional (UD) layers; unidirectional (UDW)
made from nearly unidirectional fabrics with a fill of
thin glass threads for fixation; angle-ply (0 , ___ 45 , 90 )
made from UD plies or 0 and 90 fabrics and crossply (0 , 90 ) made from weaves.

Anisotropyof composite properties


Most properties of fibre composites are of a
tensorial nature. The anisotropy is due to the fibre
arrangement, the intrinsic fibre anisotropy (carbon and
Kevlar fibres) and the fibre-matrix interfacial bond.

Fibre anisotropy
Fibre glass consists of an isotropic material with
strong covalent bonding between atoms which results
in a high strength in three dimensions. Carbon fibres
are anisotropic and have strong covalent bonding only

Properties of matrix and fibre composites


OUT,
G Pa

UT,
%

E,
G Pa

Cuc,
G Pa

t'lLS,
M Pa

dr/O'uT'b
%

0.18

2.2

0.37

4 0 at 7 7 K

140
240
130

0.8-1.3
-

0.32
0.32
-

110
90
140

85
85
-

0.7-1.0

170 a

0.32

38 a

0.07

0.1

M A T R I X A T 4.2 K
Epoxy resin Cy 2 2 1 / H y

979

FIBRE-EPOXY COMPOSITES (FIBRE CONTENT 6 0 VOL%) A T 4 . 2 K


U n i d i r e c t i o n a l II
W i t h carbon

2.0
1.5
-

1.5
0.5
-

W i t h fibre glass
(E-glass)

1.1-1.8

1.5-1.8

45

W i t h Kevlar 4 9

1.3-1.5

1.2-1.3

100

0.05-0.06

0.4-0.5

13

0.5
< 0.2

1.1
< 0.3

53
65

O.31
0.3

65
65

0.35

1.1

38

0.35

50

fibre

T 300
M 40A
A S4

65

Unidirectional 1
W i t h carbon

fibre

T 300

W i t h Kevlar 4 9

9.5

A n g l e - p l y (0 =, 4- 4 5 , 9 0 )
W i t h carbon

fibre
W i t h Kevlar 4 9

T 300
M 40A

a - at 77 K; b - fatigue endurance limit (N > 107 ) at 77 K

640

CRYOGENICS. N O V E M B E R 1 9 8 4

FIBRE ANISOTROPIES
FIBRE-GLASS

CARBON - FIBRES

KEVLAR- FIBRES
-t ..... t
N-H O=C

"-

"

C=O H-N
N-H

Oo
..0

PAN-

strong:

3-dimensional

fransversally

STRONG
Fig. I

strong

strong:
weak :

based

PITCH-

"'.

2-dimensional

l-dimensional
I (vn der Weals)
fransversol[y
medium weak
strong

BOND

[medium

WEAK

strong I dlmenslono[
weak : 2 - d i m e n s i o n a l
(van der Weals,H-bonds)
frcmsversolly weak

BOND---

Isotropic and anisotropic bonding structures of fibres

within small g r a p h i t i z e d areas which are m o r e or less


c y l i n d r i c a l l y a l i g n e d in the fibre direction ( P A N b a s e d
fibres). These areas are weakly b o n d e d by Van d e r
W a a l s forces. Therefore, their transverse strength a n d
stiffness are rather low. T h e transverse stiffness Ev_L is
h i g h e r by o n l y a factor o f two or three t h a n that o f an
epoxy resin at 4.2 K (see T a b l e s 1 a n d 2).
The a n i s o t r o p y a p p l i e s also to the t h e r m a l
properties. T h e t h e r m a l conductivity at RT is m u c h
h i g h e r in the fibre direction t h a n p e r p e n d i c u l a r to it.
At sufficiently low t e m p e r a t u r e s the intrinsic
conductivity t e n s o r o f fibres b e c o m e s m o r e isotropic
b e c a u s e only long-wave p h o n o n s are activated which
p r o p a g a t e in a s i m i l a r m o d e in the transverse a n d
l o n g i t u d i n a l directions o f g r a p h i t i z e d areas.
Kevlar fibres are even m o r e anisotropic. T h e y
consist o f stretched a r a m i d e m o l e c u l e s with strong
covalent b o n d i n g in the fibre directions only. T h e
m o l e c u l a r c h a i n s are c o n n e c t e d in two d i m e n s i o n s by
weak Van d e r W a a l s or h y d r o g e n bonds. In addition,
the Kevlar fibres have a microstructure of fibrillar
b u n d l e s with l o n g i t u d i n a l voids between them}
Accordingly, the transverse a n d s h e a r strengths a n d
stiffnesses are very low. As seen from Fig. 2 the m e a n
transverse stiffness o f Kevlar fibres is s i m i l a r to that o f
an isotropic p o l y m e r at low temperatures. The anisotropy holds also for the t h e r m a l conductivity t e n s o r at
RT. As for c a r b o n fibres the conductivity tensor at low
t e m p e r a t u r e s b e c o m e s m o r e isotropic due to long wave
phonons.
The a n i s o t r o p y of t h e r m a l e x p a n s i o n is strongly
m a r k e d for Kevlar fibres. A small negative coefficient

CRYOGENICS NOVEMBER 1984

o f t h e r m a l e x p a n s i o n exists in the fibre direction a n d a


rather large positive one p e r p e n d i c u l a r to it. This is
due to the v i b r a t i o n a l m o d e s a n d the b o n d i n g
potentials involved.
T h e structures a n d b o n d i n g forces o f the fibres
c o n s i d e r e d are shown in Fig. 1. T h e transverse a n d
l o n g i t u d i n a l m e c h a n i c a l properties are s u m m a r i z e d in
T a b l e 1. T h e m e a n transverse stiffness E v i of c a r b o n

Io ~
~

~_....

E~,
~

Kevlar-fibre

~UO
8 ~

[L v ~ ' " ~ , ~

composite

~Q~ronsverse)

E,,

" ~

Epoxyresin

Y_
cD
~

"~,,~

''1~

I
0

I
50

I
I00

[
~50

I
200

I
z50

I
300

Temperature,K
Fig. 2 Transverse stiffness of UD Kevlar fibre composites and stiffness

of the epoxy resin versus temperature

641

and Kevlar fibres was estimated from data on

Mechanical properties and thermal residual strain

transverse fibre composites and on the epoxy resin


applied. Equation (7) was modified for hexagonal
cross-sectional fibre arrangement; the transverse
modulus is a mean value composed of radial and
azimuthal fibre moduli,

Fibre-matrix interracial bond

The bond arises from covalent chemical bonds,


adhesion by weak Van der Waals bonds and interfacial
friction. The types and reactions responsible for
polymer-fibre glass bonds are more or less known and
applied in commercial composites,
For carbon fibres oxidative surface treatment has
been successful but further investigations will be
necessary. Possible types of covalent bonds between
hydroxyl groups of oxidized carbon fibre surfaces and
epoxy resins are shown in Fig. 3.3 The fibre-matrix
friction and adhesion depends on compression by
shrinkage. For composites cooled down to cryogenic
temperatures the matrix and radial fibre contraction
are decisive,
Fibre glass contracts less than polymers thus
increasing the frictional component of the interfacial
bond. For Kevlar fibres just the reverse situation holds,
The radial or transverse contraction of Van der Waals
or hydrogen bonded aramide molecules is stronger
than in most isotropic polymers thus reducing
adhesion and interfacial friction. For carbon fibreepoxy composites no great temperature dependence is
to be expected of interfacial friction.
With the present status of fibre surface treatment
and coupling agents the highest bond strength is
achieved for fibre glass-epoxy composites, followed by
carbon fibre-epoxy composites. For Kevlar fibre
compositeg only a rather poor bond strength was
achieved which reduces even more the transverse and
shear strengths. Important aspects of the interfacial
bond are reflected in the interlaminar shear strength
which is described below.

Cryogenic properties of composites


The anisotropic behaviour of fibre composites
arises from fibre and matrix dominated thermal and
mechanical components. However, several fibre
dominated properties can be determined indirectly by
the matrix behaviour. At cryogenic temperatures this is
especially true for mechanical properties which are
influenced by matrix failures or by propagating
microcracks. Their initiation is strongly enhanced by
thermal residual stresses or strains,
Bond

-~/-/0~ " ] ;-C2H 4 - N H

z +HtC-/CH

Fibre

-R~

Finish
eg
ominosilone
epoxy
resin

0 ~

QIoss y/J-O-S;-C~H4-NH -CH 2 -CH -R ~

OH
~
\0/ ~
~//~/C- 0 - CH z - CH - R ~

oxidotion
epoxy resin

OH

Fig. 3 Examplesof chemicalbonds between fibre and matrix


642

loading~

{efM} = eUTM -- {~RM}

(1)

The brackets comprise longitudinal (z), azimuthal (#)


and radial (r) components. Both the latter components
are difficult to measure separately. Therefore, approximately only longitudinal (z or II ) and transverse (J_)
components, with respect to fibre direction, are considered. For composites with anisotropic carbon and
Kevlar fibres a reasonable approximation is to assume
that the transverse fibre contraction is equal to that of
the matrix. Thus, only the longitudinal component
remains. For fibres with a negligible longitudinal
contraction (AL/LF~ 0) the thermal residual matrix
strain is given by er_M~- AL/LM within the temperature
z
range considered. For composites with isotropic low
contractive fibres, such as fibre glass, matrix shrinkage
occurs and a multiaxial stress situation in the matrix
must be considered (the equations are listed in
reference 5).
For lamination epoxy resins the ultimate
tensile strain at cryogenic temperatures 4 is:
8UTM (4.2 K) = 2.2%. The integral thermal contraction is
typically ,td-,/LM (293-4.2 K) = 1% to 1.2%. The
resulting free strain of the matrix yields the following
values:
- - composites with 60 vol% carbon or Kevlar fibres:
elM (4.2 K) = 0.7-0.9%
- - composites with 60 vol% fibre glass:
efM (4.2 K) = 0.5-0.8%.
For a matrix with prepreg resins the values of the free
strain are lower by at least a factor of two or three. As
seen from Tables l and 2 these values are insufficient
for most UD composites considered and matrix failure
occurs prior to fibre breakage.
The free strain might be little larger than
estimated. The accumulation of strain was assumed to
start at RT, although some viscoelastic relaxation may
occur even down to secondary relaxations.
When connecting UD layers as angle ply,
additional residual and thermal residual stresses,
especially shear stresses, are induced. By proper fibre
arrangement the thermal residual stresses in the
directions of external load can be minimized. This
important optimization of fibre composites has been
discussed in a previous paper? However, a rigorous
solution of problems arising from residual and thermal
residual strains can only be achieved by applying low
temperature ductile polymers which exhibit much
higher fracture strains (eg PC or PSU as mentioned

above).

Corbon //,I/C-OH -I- H.C,-C H -R ~

fibre

The main reasons for matrix failures are


brittleness and thermal residual stresses. With
decreasing temperature the polymeric matrix becomes
stiffer and stronger but also less ductile, and in
combination with low contraction of the fibres more
and more thermal residual tensile stress and strain
{eRM} is accumulated on the matrix. This reduces the
effective free stress or strain {EfM} available for external

Elastic moduli, ultimate

stress a n d s t r a i n

According to simplified assumptions the


mechanical properties of composites can be modelled

CRYOGENICS. NOVEMBER 1 9 8 4

as parallel or serial arrays of springs representing the


matrix, fibre and bond.
UDfibre composites (C) in thefibre direction ( l[ ).
For a parallel array of the fibre (F) and matrix (M)
it holds:
strain
Cell(T) ~ eM (T) -~ EF II

(2)

link between the ultimate strengths of ~rF, aM and gB.


Experimental results show that (ruTc < (rut M.
According to (6) the composite breaks at a strain lower
by a factor of (l-F) than the fracture strain of the
matrix if the transverse fibre strain is negligible. Then,
when fracturing cross-plies the matrix in the transverse
layers cracks earlier than the matrix and fibres in the
longitudinal, say load-bearing layers.

Angle-ply composites (0 , 90 ) or (0 , +_ 45 , 90).


stress
acll(Y) ~ aM(T ) ( l - F ) + OFII F

(3)

Young's modulus
EctI(T) -~ EM(T ) ( l - F ) +EFII " F

(4)

F is the fibre content by volume; polymer and


composite properties are indicated in Table 2. For a
moderate filling the composite modulus E c II is
dominated by the nearly temperature independent fibre
modulus E F II The influence of the matrix is less than
10%. Elasticity data are very similar under tension and
compression?
However, the ultimate compressive strength and
strain are always smaller than tensile properties because
buckling occurs. 7 This effect is especially serious for
transversally weak fibres, such as Kevlar. The ultimate
tensile strength rUTc and strain CUTC of U D composites
are determined mainly by the strength and content of
fibres, irrespective of matrix failure. This advantageous
behaviour applies to static uniaxial load in the fibre
direction. The situation in the more important case of
angle-plies, where cracks from neighbouring plies
influence the load bearing layers, is discussed below.
However, for several applications such as vacuum tight
vessels in cryostats, matrix break is intolerable.

Among various failure modes, such as delamination,


one failure mode will be considered which is typical of
brittle polymeric matrices. When (0 , + 45 , 90 )
composites are loaded, say in the 0 direction, crack
initiation and propagation are favoured within 90 and
- 45 layers. When cracks reach a load bearing layer
(0 ) they cut fibres as soon as the stress concentration
exceeds their strength. This is true for tranverse weak
fibres such as carbon or Kevlar fibres. Fibre glass is
rather tough and acts as a crack stopper. In Fig. 4a a
fractograph of such a carbon fibre composite shows a
plain and smooth fracture area produced at 90 and a
stepwise fracture area at 45 in the load bearing 0
layer. By contrast the rough fracture area with fibre
pullout of UD composites is shown in Fig. 4b.
Although the matrix has negligible load bearing
functions, it can reduce the strength of a carbon fibre
composite by 30% and more? This fracture mode could
be avoided by a matrix which is ductile at low
temperatures.

Fatigue characteristics
The highest fatigue endurance limit has been
found for carbon fibre UD composites; in tensile

UD fibre composites (C) perpendicular to the fibre


direction ( l ) . For a serial array of the matrix (M), fibre

o]

(F) and bond (B) it holds:

stress
Oc.L(T) ~-- aM(T) --~ OF ~ o B

(5)

strain
ecj.(T) ~ eM(T ) ( l - F ) +eF.L F

(6)

Young's modulus

1
EL(T)

,~ ( l - F )
E~(T)

F
+ ~
E~,j.(T)

(7)

For transversally stiff fibres (fibre glass. E v j_ > >


EM(T) ) the temperature dependence is controlled by
the matrix. For transversally weak fibres (carbon or
Kevlar fibres: EF ( T ) ~ EM(T)) an additional
temperature dependence arises from the fibres. The
temperature dependence of the transverse stiffness
E c is shown in Fig. 2 for U D Kevlar fibre composites.
For comparison the matrix stiffness is plotted in
addition. No great difference has been found to exist
between both stiffnesses.
The ultimate tensile strength ~rUTC and strain
CUTC of the composite are determined by the weakest

CRYOGENICS. NOVEMBER 1984

ol

!~
Fig. 4a - - Fractograph of (0"; :t: 45*; 90") angle-ply composite with
M 40A carbon fibres at 77 K. b - - Fractograph of UD carbon fibre
composites at 77 K (M 40A-carbon fibres)

643

Strength : %T and Stiffness: Ell


2.1

The value of the interlaminar shear strength


rILS ~ 170 MPa at 77 K supports the assumption that
the matrix determines ~'ILSFor UD carbon fibre composites made from
untreated fibres one gets a value of the order of
VlLS ~ 40 MPa at 77 K. Optimized oxidative treatment
and adapted coupling agents improve the ILS up to
values of VIES(77 K) ~ 110-140 MPa with H T fibres
IT 300 and AS4) and vILS(77 K) ~ 60-80 MPa with HM
fibres (M40A)? This behaviour supports the assumption
that the bond strength is the limiting factor. For HM
fibres the low transverse fibre strength might exert an

Fibre composites

6o vol*/,
2

1.8

- -

77 K

~_

i.4
j
~1

~1 b

"

#-

'~
+-45~

t~ I

t~] ~ 9 0 " )

t~

0
I00

Fatigue endurance limit ( N > I O 7) (tensile threshold)

85*/0

b5

85%~
lill
IIII s~*/,
~

5o _

zs*/,
M

65*/,
"~ 50"1,
NI ~
4o*/,T
|
,.

o
E-Gloss

roving

additional influence. For UD Kevlar fibre composites


the low value of VlLS(77 K) ~ 38 MPa indicates failure
by debonding and probably also by the fibres.

T300

M40A

Kevlor49

Corbonfibres

Fig. 5 Tensile strength GUT, stiffness E II and fatigue endurance limit a / % T


of UD and (0"; :t: 45*, 90*) fibre composites at 77 K

Thermalconductivity
A survey of thermal conductivity of UD fibre
composites is given in Fig. 6. Fibre-glass and Kevlar
fibre composites do not show a great temperature
dependence of thermal conductivities) H3 their values
being five to ten times higher than that of the epoxy
matrix. By contrast, carbon fibre composites show a
very large temperature dependence due to freezing out
of electron thermal fibre conductivity at low
temperatures.
In Fig. 7 the thermal conductivity of carbon fibre
composites is shown with different types and
arrangements of fibres. At RT all UD composites with
HM fibres (M 40A) show a higher conductivity

threshold tests at 77 K the value is 85%. The values of


other UD and angle ply composites are shown in
Fig, 5. Tests under compressive fatigue loading yield lower
values because of microbuckling, especially for fibres
which are weak in the transverse direction. The data
are very scarce. Some information on fibre glass
composites has been published?
One important question is the degradation of
properties in the course of fatigue cycling. For fibre
glassandcarbonfibrecross-pliesitwasshownthat
stiffness does not degrade by more than 15% under
uniaxial tensile fatigue conditons at 77 K until fracture
occurs. Under biaxial fatigue cycling with combined
torsion and tension a much higher and cumulative
degradation has been found. 1

(so vol %)

For shear loading of composites the decisive


limiting parameter is the ILS. If this value is exceeded,
delamination of composite layers occurs. Fibre
breakage occurs at loads defined by bending strength.
The ILS measurements are performed in short beam
tests. The values of ILS are determined mainly by
failures of the matrix, interfacial bond or fibres.
The question of the upper limit of ILS arises and
another question is which components constitute the
limitations of the composites considered. The low
temperature shear strength of epoxy resins (and most
polymers) lies between 150 and 200 MPa at T ~ 77 K
and constitutes the upper limit of composites containing epoxy resins. For fibre glass composites the
high, nearly isotropic fibre strength and very likely the
bond strengths do not constitute the limiting factors,

644

19140A

/ - ~
/

Io 1~

(Steel)

/,/"

!/'/"

poxy . . . ~ . - -

x,, __

'l[]/_....-~-I
......_~_
.......-~

Interlaminarshearstrength(ILS)

?11.~.__

Composites

~ io-I I

__ -- --

T 300

Kevlor/epoxy
Gloss/epoxy
Epoxy

162

i63

I I

Fig. 6

I I

I I

too

zoo
Temperoture,K

I ~ ~ i i

i ~ i

3o0

Survey of thermal conductivities of UD fibre composites versus

temperature

CRYOGENICS.

NOVEMBER

1 984

Carbon fibres/epoxy
60 vol %

150
M40A II

~E

xlO -4

M40A_+450

ooo

///,""
/ f
///~"

~
/
f

T 3 0 0 II
TSO0 0", +-45",90
T:50045

Kevlar fibre-composite
Fibre content 60 vol%

~-E.~ox~yr
-

oo i
50

- ~'~,~

~,,~ransver se

m ~0

~" "~
~" "... "-.
I

-,o

~///~'///
~ / /

I0- I

Epoxy resin
CY221 HY979

xtO"5
-I00

--

6
4
5

16 2

Fig. 7

I I I I II I II I
20 40 60 80 IO0

150

200

250

500

7", K
Thermal conductivity of different carbon fibre composites versus

temperature

compared to those with HT fibres. The electric


conductivity of HM fibres at RT is about 2.5 times
higher than that of HT fibres. This is due to larger
graphitized areas of HM fibres,
At very low temperatures this difference vanishes
because wavelengths which are characteristic of that
temperature range (dominant phonon wavelength) are
too large tbr resolving the microstructures of fibres. A
similar consideration holds for the longitudinal and
transverse components of the fibre conductivity tensor,
At RT it is rather anisotropic and gets more isotropic
at low temperatures due to the increase of the
dominant phonon wavelengths. In addition the carbon
fibre conductivity decreases with decreasing
temperatures and gets comparable to or lower than that
of the epoxy matrix.
As seen from Fig. 7 the different types of carbon
fibre composites considered vary by a factor of 50 at
RT. At 7 K this factor reduces to 1.5. For the thermal
conductivity transverse to the fibre direction a thermal
boundary resistance by phonon mismatch (Kapitza
resistance) may be relevant at very low temperatures.
This means that at low temperatures the thermal
conductivity of carbon fibre composites is rather small
and nearly independent of fibre type and arrangement.
Thus thermal conductivity imposes no major restrictions if one wants to optimize other composite
properties at low temperatures.
Thermal

expansion

For several applications in cryogenic temperature


technology low contractive materials are of advantage,
Due to the anisotropic tensor of thermal expansion this
is not possible for all fibre composites in all directions.
Fibre glass has a low coefficient of thermal
expansion (atr ~ 4.8 10-n K-1) which is nearly isotropic
in all directions. Thus, the contraction of those
composites simultaneously can be made small in all
three directions,
Carbon and Kevlar fibres exhibit a very small or
even negative coefficient of thermal expansion only in

CRYOGENICS.

NOVEMBER

1984

I ~
I00

I I
200

I
300

Tempereture, K
Fig. 8 Longitudinal and transverse integral thermal expansion for UD
Kevlar fibre composites versus temperature

the fibre directions due to stretching vibrations in


covalent bonding potentials. The negative expansion of
Kevlar fibres is assumed to arise from bending
vibrations perpendicular to the aligned molecular
chains of fibres. In contrast their transverse expansion
is large and larger than that of most epoxy resins. This
is, due to vibrations in the weak asymmetric Van der
Waals or hydrogen bond potentials. Thus, a low
thermal expansion is only available in the fibre
direction. This is demonstrated in Fig. 8 for the
thermal expansion in the longitudinal and transverse
directions of Kevlar fibre UD composites./3 For carbon
fibre composites similar results apply.
In angle-plies a medium low thermal contraction
can be achieved in plane by means of residual stresses
between plies. However a large thermal contraction
exists in the direction perpendicular to the laminates.
The coefficient of thermal expansion of UD
composites atc II can be calculated 5 from those of the
fibres o~F II and the matrix o~M by
C~M --~rll
Sell = erll +

(8)
1 + 1.1 F/(1-1.1/7) (Evlp/EM)

where F is the fibre content per volume: E M and


EF II are the moduli of matrix and fibre, respectively.
For calculating thermal expansion of angle-plies see a
previous publication?
The thermal expansion can be varied by the fibre
content and arrangement or by different thicknesses of
angle-plies.
Experimental values of integral thermal
expansion
4.2

AL/L =

e c dT

293
on angle-ply fibre glass composites are shown in Fig. 9.
The parameter of the curves is the fibre angle __. to
related to the direction of measurement) 4

645

Fibre gloss/epoxy
70 -',c5_ vol%

Fibre gloss/epoxy
70 -+ 3 vol %

I~'~x
= 90

Directionof
measurement

~_

:500

+<~

Direction of
+~_~_+~
measurement

(d

= 60 =

300
(

Thicknesses of plies
,

Is=tso./t
1

;~0

200

0)=45

200

--~=0.1

"~

a3

~t3=0.5

oJ= 450
~

0)=30 `

to=(

I00 --

o I

t
I00

t
200

300

Temperature, K

Fig. 9 Integral thermal expansion of balanced angle-ply fibre glass


composites versus temperature. The parameter is the fibre angles to
I

In Fig. 10 the values of composites with three


fibre directions (___to; 90 ) are shown. Additional
parameters are the relative thicknesses of angle-plies. Is
Both figures are examples demonstrating the great
v a r i a b i l i t y o f fibre c o m p o s i t e s . I n m a n y c a s e s t h e s a m e
contraction behaviour can be achieved by different
fibre arrangements. This opens up the possibility of
optimizing additional requirements, eg thermal or
mechanical properties.

Dependence on production technique and


fibre

lay-up

The data summarized in Table 2 are valid for a


wet lay-up technique and for an exact fibre alignment
of each ply. Composites made from fabrics yield values
of fracture strength lower by more than 30%. This
arises from the wave shaped misalignment and the
transverse compression exerted by fill and warp. The
latter is especially of influence for transverse brittle
fibres such as carbon and Kevlar fibres,
The interlaminar shear strength tiES is sensitive
to the quality of lamination. Two examples are given
for UD composites with carbon T 300 fibres
tiES(77 K) ~ 110-140 MPa (wet lay-up technique)
tiES(77 K ) ~

80-95 MPa (prepreg technique)

An appropriate compromise is a matter of material


quality and economy,

646

I
I00

I
200

IkL
300

Temperoture,K
Fig. 10 Integral thermal expansion of balanced angle-ply fibre glass
composites versus temperature. The parameters are the fibre angles to

and the relative thicknesses t"a of the 90* ply

Conclusions
An important advantage of polymeric fibre
composites is their great variability in matching
mechanical and thermal properties. In cryogenic
applications the brittleness of the epoxy resin matrix
reduces the mechanical strength, especially under
multiaxial load conditions. The study of degradations
of composite properties under fatigue cycling is a
subject of further investigations. For future developments a substitution of brittle epoxy resins by special
low-temperature ductile thermoplastics would help to
reduce thermal residual stresses and stress concentrations within fibre composites. This requirement
becomes more stringent with the improvement of fibre
properties. Carbon fibres with ultimate tensile strains
of 1.8% will be available in the near future. An
additional advantage is that thermoplastics can be
welded by solvent agents or by heat. This improves the
quality of vacuum tight connections of members made
from thermoplastic fibre composites. After curing

CRYOGENICS. NOVEMBER 1984

epoxy resins can only be glued. This is m a i n l y a n


a d h e s i o n by weak Van der Waals forces,
Each type of fibre composite has its specific
merits. A n appropriate c o m b i n a t i o n of different fibre
types in hybride composites opens up a large field for
o p t i m i z i n g composite properties.

Authors
The authors are from the K e r n f o r s c h u n g s z e n t r u m
Karlsruhe, Institut ftir Reaktorbauelemente, D7500,
Karlsruhe, F R G . Paper received J u n e 1984.
Most of the samples were prepared by
Messerschmit-B61kow-Blohm, M u n i c h . T h e
c o l l a b o r a t i o n of B. Vogeley a n d Dr. Weiss is greatly
acknowledged.

References
1
2

Hartwig,G. Proc lnt Cryogenic Materials Conf, Eds K.


Tachikawa and A. Clark. Butterworths (1982) 495
Preston, J. Aramid Fibres Encyl. Chem. Technol. 3, Wiley,
New York, 3rd Edn. (1978)

CRYOGENICS. NOVEMBER 1984

Fitzer, E., Weiss, R. Processing and uses of carbon fibre


reinforced plastics, VDI-Verlag, Diisseldorf(1981) 45
Hartwig, G., Jaeger, H., Knaak, S. Nonmetallic Materials
and Composites at Low Temperature II1; Eds. G. Hartwig,
D. Evans: Plenum Press New York (1985)(in print)
4 Hartwig,G. Nonmetallic Materials and Composites at Low
Temperatures I: Eds. A, Clark, ILP. Reed; G. Hartwig; Plenum
Press, New York, (1978) 33-50
5 Schneider,W. Kunststoffe 61 (1971) 273
6 l~sen, M.B. Adv Cryo Eng (Materials) 28 (1982) 165-177
7 Chamis,C.C. Composite Materials 5 Ed. L.J. Broutmann,
Academic Press, New York (1974) 121
8 Hartwig,G. Adv Cryo Eng (Materials) 28 (1982) 179-189
9 Nishijima,S., Keta, S., Ol~da, T. Nonmetallic Materials and
Composites at Low Temperatures; Eds: G. Hartwig, D.
Evans; Plenum Press New York, (1982) 139-149
10 Wang,S.S., Chim, E.S,M. Adv C~'o Eng (Materiala9 28 (1982)
201-210
11 Rosenberg,H. Nonmetallic Materials and Composites at
Low Temperatures I1: Eds: G. Hartwig~ D. Evans: Plenum
Press New York, (1982) 181-196
12 Hust, H.G. Cryogenics 15 (1975) 126 and Kasen, M.B.
Cryogenics 21 (1981) 323
13 Hartwig,G. Tieftemperaturtechnologie, Ed. F.X. Eder,
VDI-Verlag, Diisseldorf, (1981) 177
14 Hartwig,G., Puck, A., Weiss, W. Kunststoffe 64 (1974) 32
15 Bauder,P., Hartwig, G,, Sehindler, R. J Mat Sci 22 (1976) 225

647

S-ar putea să vă placă și