Thermochemical model for chemically

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Int. J. Engng Sci. Vol. 35, No. 2, pp.

113-128, 1997
Pergamon Copyright ~) 1997 Elsevier Science Ltd
PII: S0020-7225(96)00072-9 Printed in Great Britain. All rights reserved
0020-7225/97 $17.00+ 0.00

A THERMOMECHANICAL MODEL FOR CHEMICALLY


DECOMPOSING COMPOSITES--I. THEORY

YINAN WU and NORIKO KATSUBE


Department of Engineering Mechanics, The Ohio State University, Columbus, OH 43210, U.S.A.

(Communicated by J. T. ODEN)

Abstract--In this work, a thermostructural model for chemically decomposing composites with
moisture is obtained based on the mixture theory by Green and Naghdi. Balance laws of mass, linear
momentum, and energy in the mixture theory are expressed in terms of physically measurable
quantities. The formulation includes moisture vaporization, thermochemical decomposition, thermoo
mechanical deformation, gas flow, and heat transfer. It is shown that the usual form of equilibrium
equation and Darcy's equation, widely used in literature, need modification when the phase change
occurs. Moisture vaporization is incorporated in mass and energy balance, and vaporization rate at
each point of a continuum can be theoretically predicted. Copyright © 1997 Elsevier Science Ltd

INTRODUCTION

The thermomechanical response of chemically decomposing carbon-phenolic composite


materials has been of great interest in the space industry. These composites are used as heat
liners for the solid rocket motor nozzles of space shuttles. In actual flights, these composites
protect the metal of nozzles by forming a char layer through thermochemical decomposition,
which is energy absorbing. In ideal conditions, the char layer gradually advances from the inner
surface of the nozzle toward the interface between metal and composite. The thickness of the
composites is designed so that the flights are over before the decomposing region reaches the
metal.
In reality, however, two types of premature failures have been observed. In one type of
failure known as ply lift, layers of woven carbon fiber have been observed to separate. In
another type of failure known as pocketing, fracture of woven carbon layers has been observed.
The composites become porous upon thermochemical decomposition. The decomposing gas is
often trapped inside the composites and exerts high pore pressure. This high pore pressure is
considered to be the main cause of these failures. If moisture is contained inside the
composites, vaporization of the moisture contributes to the formation of high pore pressure.
Mathematical models and numerical methods have been developed to analyse the thermal
response of decomposing materials. For example, Henderson and Wiecek [1] have proposed a
one-dimensional mathematical model for analysing the thermal response of decomposing
polymeric composites. Temperature, pore pressure, and solid mass degradation are calculated
based on the conservation of energy, gas mass continuity, and chemical kinetic equations. In
characterizing the material deformation, an empirical model based on Buch [2] is employed to
curve fit the thermal expansion data. This approach is effective in simulating the thermochemi-
cal expansion behavior without mechanical load, but is inadequate to predict the structural
response of the composite.
More recently, various pioneering works have been done to mathematically model the
coupled physical phenomena and numerically simulate experiments. For example, in Sullivan
[3] and Sullivan and Salamon [4,5], the poroelasticity theory [6] and thermochemical
decomposition analysis [7] are combined. Their FEM code is for 2-D plane problems, and
laboratory experiments such as free thermal expansion tests and restrained thermal growth
tests are numerically simulated. This method is applied to analyse the material response of a
glass-phenolic composite subjected to thermochemical decomposition [8]. Salamon and Lee
113
114 Y. wu and N. KATSUBE

[9] have further extended the method and formulated an axisymmetric finite element code. This
model is also applied to analyse a thermal insulation liner in the cowl region of a typical solid
rocket nozzle to investigate the mechanism for plylift. In addition, in Sullivan [10, 11] the effect
of moisture on thermal stress is modeled based on a mixture theory approach. The result from
his model is in close agreement with experimental measurements. The code developed by
Kuhlmann [12] is for 1-D axisymmetric problems. It includes the rate of moisture vaporization
in the form of phenomenological equations based on empirical data. A finite difference code
for 1-D thin plate has been developed by McManus [13] and McManus and Springer [14, 15]
and verified against their experimental data. Their work includes modeling of material ablation
and mechanical damage to simulate the delamination of fiber plies. By assuming that gas
storage is small compared to gas diffusion, temperature and pore pressure are solved from the
simplified forms of heat transfer and gas flow equations. The effect of material deformation on
gas flow is not considered. Without considering moisture effect, Weiler [16] has obtained fully
coupled governing equations for deformation, gas flow, and heat transfer solutions based on a
phenomenological approach. In Wu and Katsube [17, 18], the attention is focused on the
thermomechanical response of the composite without considering gas flow. The governing
equations used by these researchers are not exactly the same. Comparison among these works
and clarification as to which equations should be used are much needed.
In this work, the mixture theory by Green and Naghdi [19] is applied to the decomposing
composites with moisture from the point of view that the form of governing equations must be
usable. A thermostructural model is established by expressing balance laws in terms of
physically measurable quantities with some simplifications and approximations. The model
includes moisture vaporization, poroelastic deformation, thermochemical decomposition, heat
transfer, and gas flow through porous materials. The general nature of the resulting formulation
is demonstrated by comparing it with formulations of others.

PRELIMINARIES

The composite consists of a porous solid material with gas inside the pores. Deformation of
the composite is assumed to be infinitesimal and elastic while solid mass loss due to
thermochemical decomposition can be finite. The composite contains moisture at room
temperature. As temperature increases, moisture desorbs and becomes vapor. In addition, the
virgin solid phase experiences a series of chemical reactions, and volatile substances are
generated due to decomposition. Vapor and volatile substances mix together inside the pores.
This mixture comprises the gas within the composite.
The basic framework of mixture theory by Green and Naghdi [19] will be employed. Each
point of the mixture continuum is occupied simultaneously by two constituents. If gas is
removed from a gas-filled solid material, a porous solid material is recovered. Similarly, if solid
material is fictitiously removed from the gas-filled solid material, a porous gas material is
recovered. The porous solid material and the porous gas material are idealized as an equivalent
homogeneous solid material and an equivalent homogeneous gas material, respectively.
Therefore, two different constituents in the mixture theory physically represent these
equivalent homogeneous solid and gas materials, denoted by sl and s2, respectively. Since
moisture is absorbed inside the solid, moisture itself will not be considered as a separate
continuum. The position of each continuum at time t is described as

xi = x i ( X l , X2, X3, t) Yi = Y,(Y1, Y2, Y3, t), (1)


where xi and Yi are the positions of sl and s 2 at current configuration, respectively; X A and YA
Thermomechanical model for chemicallydecomposingcomposites 115

are the positions of sl and s2 at reference configuration. Particles of these two continua under
consideration are assumed to occupy the same position at time t so that

Xi : Yi. (2)
(1) (2)
Let D / D t and D / D t represent the material time differentiation of sl and s2, respectively.
They are related to the spatial time differentiation a/at in the spatial frame of xi as follows:
(1) (2)
D 0 0 D 0 0
-- = -- + v,~ ') (3)
Dt Ot OXk' Dt Ot aXk

where the velocities of Sa and s2, v}1) and v}2), are defined by
(1) (2)
Oxi xA Dxi V~2) = Oyi = Oyi
v!l) = -~t = Dt' ~ - vA Dt (4)

Equation (3) also gives the following relation between these two material time differentiations:
(2) (1)
(O) O 0
= -- + v~-- (5)
Dt Dt OXk

where the relative velocity vi is given by


Vi = V~2) -- V! 1). (6)

CONSERVATION LAWS

Mass balance
G r e e n and Naghdi [19] have postulated an energy balance equation. The use of invariance
conditions under superposed rigid body motion leads to the mass continuity equation

MsL + Mg = 0, (7)
where Mse and Mg, respectively, are defined by
(1) (2)
MsL -- DOsE + OsLV~',~, Mg = DOg + . ,,(z) (8a, b)
Dt Dt ~'g--k,k

and physically represent the rates of solid mass generation and gas mass generation in a unit
volume of the mixture. The bulk densities of the wet solid phase and gas in the composite are
represented by PsL and pg, respectively. They are related to the real densities of the wet solid
phase and gas, denoted by P s L and/~g, as follows:

PsL ---- (1 -- ~ ) P s L , p g ~---~t)pg, (9a, b)


where ¢p is the porosity of the composite. The mass balance equation for the wet solid, equation
(8a), can be separated into two parts. Introducing the bulk densities of dry solid and that of
liquid water, denoted by ps and PL, respectively, we have

RsL = Ps + RL. (10)


Introducing the rate of water mass loss due to vaporization and the rate of mass loss of dry
solid due to thermochemical decomposition in a unit volume of mixture, denoted by ME and
Ms, respectively, we have
Ms + ME = MsL (11)
116 Y. WU and N. K A T S U B E

with
M~, =/5~ + ~'~,,0)
k.k~ ML = PL + P L ~,,o)
k,k (12)
(1)

where a dot over a quantity is used to express D / D t for simplicity.


With the help of equations (5) and (6), the mass balance equation for the gas is obtained
from equations (7) and (8b), as in equation (13):

tSg + ~,g~v,,~"(')+ [pgVk],, = --MsL. (13)


Eliminating pg from equations (9b) and (13), the mass balance equation for the gas can be
expressed as
Pg ~, (I) 1 MsL
q- ~ ~ - q- ~p'Vk, k Jr- -- [~gPgYk],k -- (14)
Pg Pg Pg

Both thermomechanical deformation and thermochemical decomposition influence the value


of porosity. In order to examine this p h e n o m e n o n in detail, we shift our attention from the
mixture theory to micro-mechanical analysis. Let us consider a representative volume element
of the composite material. The representative volume element should be large enough to
contain the detail of the microstructure (several layers and pores), but small enough so that its
average response is representative of the local response of the composite material idealized as a
continuum. The total volume of the representative volume element consists of the solid volume
V s and the pore volume V p as in equation (15). The porosity ~p is expressed by equation (16).
V = VS+ V p, (15)
Vp
= ~-. (16)

In reality, the mass loss due to decomposition and deformation occurs concurrently.
However, as in Wu and Katsube [17], if we assume that the thermochemical decomposition is
not influenced by the stress state of the composite, it is possible to separate thermochemical
decomposition analysis from thermomechanical deformation analysis. For this purpose, we
consider a state where only the thermochemical decomposition process has occurred. Let
subscript 1 indicate this state. The total volume of the composite, solid material volume, pore
volume, and porosity at state 1 are denoted by I/1, V~, V p, and ~Pl, respectively. They also
satisfy the modified equations corresponding to equations (15) and (16). Based on this
separation, V, V s and V p can be expressed by
V = g 1 -1- AV, V s = V] q- A V s, g p = V p + AV p, (17)

where A represents the change of volumetric quantities due to thermomechanical deformation


at time t. Assuming small thermomechanical deformation, AV and AV s are respectively related
to the volumetric strain of the porous material ekk and that of the solid e~,k as follows:

A V = ekkV, A V ~ = e~,,,V s. (18)

Before the material is subjected to thermochemical decomposition and thermomechanical


load, the composite is virgin. This initial overall volume and pore volume are denoted by Vo
and Vop. The initial (virgin) porosity of the composite, denoted by ¢o, is defined by the
corresponding equation of (16). After all the thermochemical decomposition is completed, the
composite becomes char. The porosity at this state, denoted by ~f, is normally measured
without any mechanical load and at room temperature. Let VF and Vf represent the pore
volume and total volume of char composite measured without any mechanical load and at room
temperature. ~of, VF and Vf also satisfy the same relation as described by equation (16).
Due to the change in chemical structure of the polymeric material in thermochemical
decomposition, Vf may not be equal to Vo even though they are both measured without any
Thermomechanicalmodelfor chemicallydecomposingcomposites 117

mechanical load and at room temperature. When Vf is less than Vo, the composite exhibits a
volumetric shrinkage. In order to describe the volumetric shrinkage during the process of
thermochemical decomposition, we introduce the measure of shrinkage, if, defined by

v~-Vo
~"= (19)
Vo
For the representative element inside the composite, thermochemical decomposition takes
place on the solid-gas interface. It is assumed that the total volume of the representative
element, V1 in equation (19), is not affected by the mass removal resulting from thermochemi-
cal decomposition. The shrinkage at char, denoted by ~'f, can be calculated from the measured
quantities at virgin and at char. Since it is difficult to distinguish between thermochemical
decomposition and deformation, measurement of shrinkage during the process is not available
at present. It may be assumed to be linearly dependent on the degree of char, as proposed by
Wu and Katsube [18].
The overall volume can be obtained by inserting equation (18a) into equation (17a) and
eliminating 111 from the resulting equation and equation (19):

1+~"
V- - - Vo. (20)
1 - ekk
Similarly, the volume of the solid phase can be rewritten as in equation (21) by inserting
equation (18b) into equation (17b) and then eliminating V] by using equations (15) and (16) at
state 1 and equation (19):

VS=( 1 1+~ (21)


- ¢1) 1 - e~----7 Vo.

The porosity can be expressed in terms of overall strain and solid strain by eliminating V p from
equations (15) and (16) and inserting equations (20) and (21) into the resulting equation. After
linearizing the expression using the small deformation assumptions, we have

~ = ¢, + (1 - ~,)(ekk --e~k) (22)


where qh will be later determined by equation (72).

Conservation o f m o m e n t u m
The equations of motion for continua sl and s2 can also be obtained from the invariance
conditions under superposed uniform rigid body motion and energy balance [19]. The
conservation of linear momentum for the mixture is given by

(O'ji -~ ]'~ji),j -~-PsLFi Dr.pgGi = PsLf~ + Pg& + MsL V}') + Mgv~ 2) (23)
where F~ and Gi, respectively, are the body forces per unit mass of s~ and s2. The partial stresses
of s~ and s2 are denoted by o-0 and zrij, respectively. The accelerations of sl and s2 are denoted
by f and &, respectively. The total stress of the mixture is defined by
r,j = o-,j + & # (24)
It can be shown that the total stress is symmetric, such that
r,j = lrji. (25)
The partial stresses can be decomposed into the symmetric parts and the skew-symmetric
parts [20]. The symmetric part of the partial stress of &, denoted by o'(im is related to the solid
stress ~,j by
or(q) = (1 - ¢)r~j. (26)
118 Y. WU and N. KATSUBE

Similarly, the symmetric part of the partial stress of $2, denoted by rq0), is related to the
hydrostatic gas pressure p by
]r,(ij) = -- qop6i] , (27)
where p is defined positive for pressure. In equation (24), the skew-symmetric parts of partial
stresses cancel each other out. Inserting equations (26) and (27) into equation (24), the total
stress can be expressed in terms of solid stress and pore pressure by
r,~ = (1 - ~ ) r ~ j - ~op6ii. (28)
Substituting equations (7) and (24) into equation (23), the equation of motion for the
mixture with mass interchange between sl and s2 is obtained as follows:
Vii,i + PsLF/ + pgGi = P s L f + Pggi -- MsL Vi • (29)
The third term on the right hand side of equation (29) represents the effect of phase change
between continua Sa and s2 on the conservation of linear momentum for the mixture. When
both inertia and body force are ignored, the equilibrium equation for the mixture can be
simplified as
"~ji,j ~--" -- MsL vi. (30)
If there is no phase change or if the relative velocity is zero, equation (30) reduces to the usual
homogeneous form of the equilibrium equation. Equation (30) indicates that the form of the
equilibrium equation may need modification for problems involving phase change in the
material. The modification term is determined by the rate of phase change and the velocity of
gas relative to the solid.
In Green and Naghdi [19], the conservation of linear momentum for continuum s2 is given by

TCji,j 3c fl2i q- p g G i = Pggi + 2


1 MsL v (1)
i + 21_Mg v t(2) (31)

where
1 1 1
~i = 2 (O'ki -- 7"~ki),k "~- 2 OsL(Fi -- f ) -- 2 O g ( G i - g,). (32)

Based on Crochet and Naghdi [21] and Katsube and Carroll [20], anisotropic linear constitutive
relations are given by
A(2)+ SqklAk I + Ti]k Vk
rco = - ~pSij + L~ijklV,'kl
i
zi = - Q i j v j + RijkA]k (33)
where dbz) are the components of the strain rate tensor of continuum sz; Aij are the components
of the vorticity tensor of continuum sl relative to continuum s2; and Lokt, Sijkt, Tijk, Qij and Rijk
represent the components of the material constant tensors. Inserting equation (33) into
equation (31), we obtain
1 1
--(qoP),i + Lqktd~2!j + SqktAktd + TiykVkd -- Qiyv/ + g q k A j k = Pg(gi -- G i ) + "~ MsLV}1) + ~ Mgv} 2). (34)

If we ignore the gradients of velocity and porosity, body force and inertia terms, equation (34)
reduces to
[ 1
--P,i = 1q~ Q,j _ 2 M~c6,i vj
] (35)

with the use of equations (6) and (7). It has been shown [20] that when MsL = 0, conservation of
linear momentum of continuum s2 reduces to Darcy's law of the form
__K_~
v, - p j, (36)
Thermomechanical model for chemically decomposing composites 119

w h e r e / , is the viscosity of the gas. The components of permeability tensor represented by Kit
are related to Q~j by
Kit = lz~02Q/71. (37)
With the help of equation (37), equation (35) with non-zero MsL can be rewritten in the form of
Darcy's law as
AikKkj
vi = - pj (38)
/x~0
where
1 l 1

Aij : [6ij MsLK,j] . (39)


Equation (38) can be considered as a generalized form of Darcy's law. If there is no mass
interchange, Aq is equal to the identity matrix, and the usual form of Darcy's law is recovered.
Since the phase change is important in thermochemically decomposing materials, equation (38)
will be used instead of equation (36) to express the conservation of linear momentum of the
gas.

Conservation o f energy
Since the composite is exposed to an extreme thermochemical environment, heat transfer is
assumed to be dominant compared to kinetic energy and mechanical work in the energy
balance equation. If the contribution from the interaction between the constituents is ignored,
the total internal energy of the mixture can be assumed to be the sum of the internal energy of
each constituent. Let A be an arbitrary fixed closed surface enclosing a volume V, and let nk be
the outward unit normal to A. If the effect of mechanical power on the energy balance is
ignored, the global form of the energy equation postulated in Green and Naghdi [19] is reduced
to

L
a'-tt [psUs-i-pLUL-bpgUg]dV-F [Hk(psV~I)Us-FpLv~I)UL-FPgV~2)Ug)]dA

= L [Msfic)+ M L ~ ' ) l d V - L n k q k d A , (40)

where the heat flux is denoted by ql; the internal energies per unit mass of solid, water, and gas,
respectively, are denoted by Us, UL, and Ug; the reaction heat per unit solid mass loss due to
chemical decomposition is denoted by r--(c); vaporization heat for the phase change of unit mass
of water is denoted by Y(v). Both r--(c) and Y(~) are defined positive for exothermic processes at
reference (room) temperature To and pressure Po. The internal energy is assumed to be
approximated by enthalpy h as in Weiler [16]. The enthalpies of solid phase and liquid water
are defined as

h = Cp dT, (41)

where T is the current temperature and Cp is the specific heat. The enthalpy of gas depends on
both temperature and pressure. In the special case of ideal gas, its enthalpy can be expressed in
the form of equation (41). From equation (41), we have
l.f = CpT, U,k = CpT, k. (42)
The local form of energy equation (40) is obtained as
0
at [osUs + OLUL + p,U,] + (psv~')Us + OLV~')UL + pgV~Z)Ug),k = Msr-(c) + M L ~ ~') - qk.k. (43)
120 Y. WU and N. K A T S U B E

The first term on the left hand side of equation (43) is the rate of energy storage increase per
unit volume of a mixture fixed in space. The second term indicates the rate of heat convection
accompanied by the mass flow from this volume. The right hand side of equation (43)
represents the rate of heat supply due to chemical reaction, vaporization, and heat conduction.
Rewriting equation (43) using equation (3a), and substituting equations (12) and (13) into
the resulting equation, the heat transfer equation can be expressed as
ps/]s + pL/]L + pgOg + pgVkUg,k + q~,k = M, r(c) + ME r(~) (44)
where
r(C) = pc) + Ug - Us, r (v) = r-(Y) + Ug - UL. (45)
r (¢) and r ~) physically represent the heat of chemical reaction and that of vaporization at
temperature T, respectively. Further use of equations (9b) and (42) in equation (44) leads to
the following form of the heat transfer equation.
[psCps + p L C p L + ~0/~gCpg] 7" "1- ~OpgCpgVkT, k -1- qk,k = Ms r(~) + ML r(e). (46)

MATERIAL RESPONSE

Thermomechanical d e f o r m a t i o n o f a gas-filled p o r o u s c o m p o s i t e material


For the elastic response of a gas-filled porous material, the stress and strain are related by
[22, 231
~'ij = Miikt[ek,- ~ k , ( T -- To)] (47)
where Lj is an effective stress tensor, defined by
~ij = "l~ij J¢- aijp (48)
and
Otij = ~ij -- CS,mklMknj • (49)
The solid (without pores) is also assumed to be linearly elastic:
rsj = M~jk,[e~,- ~ k , ( T - L)]. (50)

In equations (47)-(50), M~j~z and M~jk~are the components of the elastic modulus tensor of a
solid (without pores) and those of a porous solid material, respectively. The corresponding
compliance components are C~j~z and C~jkt. The pore pressure and thermal expansion coefficient
tensor are denoted by p and/3ij, respectively.
Eliminating ~/, r;j and r~} from equations (28), (47), (48) and (50), the volumetric strain of
the solid phase (without pores) can be expressed in terms of pore pressure p, temperature T,
and the overall strain e~i:
1
eskk = 1--~-~ [(Sij -- aij)eij + (¢PSij - aij)(6kl -- O~kz)CijklP -- (~06ij -- aij)t~ij(T - T,,)]. (51)

The thermochemical shrinkage phenomenon causes negative normal strains. In order to


calculate the actual strain e~j of the composite material, the contribution due to thermochemical
shrinkage needs to be added to the overall strain due to deformation:
eij = eij + ~ij, (52)
where ffij is the diagonal tensor of shrinkage strain. This tensor is related to volumetric
shrinkage by
~kk = ~. (53)
By replacing q~ by ~01 in equation (51) under the assumption of small deformation and inserting
Thermomechanical model for chemically decomposing composites 121

the resulting equation into equation (22) using equation (52), the porosity at time t can be
obtained:
= ~1 -[- (Olij -- ~1 ~ij)[eij -- ~ij -- [3ij(T - To) + (~kt -- akt)Cok,P]. (54)

Vaporization of moisture
Due to the heat supplied to the material, the moisture vaporizes and becomes steam. In this
vaporization process, water and steam coexist inside the pores, and moisture is at saturation
status. During the saturation period, pressure and temperature are not independent. Their
relationship is specified by the phase diagram of water, and may be symbolically expressed as

f ( p , T) = O. (55)

This relation is tabulated and available in the steam table.


We introduce a parameter 3' to represent the amount of moisture inside a wet composite at
time t. It is defined as the ratio of water mass at time t over the total mass of the composite at
virgin:
mWater
3' - (56)
poVo
where m water and po respectively represent the mass of liquid water in the representative
volume V and the density of the composite at virgin. Since ME represents the rate of water mass
loss in a unit volume of mixture, we obtain
/~lwater
ME -- (57)
V

Since thermomechanical deformation is assumed to be small, V/Vo can be approximated by


1 + e~k. Inserting equation (56) into equation (57), ME in equation (57) can be related to 3' as

~0
ME - 1- +
- .
ekk (58)

Thermochemical decomposition
Degree of conversion. In general, there may be a series of chemical reactions occurring
simultaneously during the degradation process. Each reaction can be separately modeled by an
Arrhenius equation. Let momatrix and m~aatrix denote the mass of the original matrix in the
representative volume element and that of the degrading matrix for the kth reaction. Let Wok
and Wf~, respectively, represent the fractional mass of m matrix during the process and that left as
the solid residue at the final stage. Therefore, the value of m ~ aatrix changes from Wokmmatrix a t
the beginning t o Wfkmmatrix at the completion of the kth reaction. The degree of conversion for
this reaction, denoted by ck, is defined as the mass loss that is to occur before the completion of
the reaction over the total mass loss for this reaction as follows:
m ~ atrix -- Wfkmmatrix
Ck -- Wokmmatrix _ Wfkmmatrix . (59)

The value of Ck varies from one at the beginning to zero at the completion of the kth reaction.
The Arrhenius kinetic reaction ecluation is given by

~ _ ~ = _ _ Z o k C ~ k e x p ( __ E a k )
RT/ ' (60)

where A,k, E,k and nk are the reaction rate constant, the activation energy, and the order of
reaction, respectively. The universal gas constant and absolute temperature are denoted by R
and T, respectively.
122 Y. W U and N. K A T S U B E

As in the degree of conversion for each reaction [equation (59)], the degree of conversion for
the entire degradation process of matrix, denoted by c, is defined by
mmatrix __ m ~ natrix
c----
mo
matrix m~natrix ' (61)
where
1
mmatrix = E m ~ atrix, (62)
k=l
/
m ~ natrix E matrix
= Wfkmo • (63)
k=l

The number of reactions is represented by l in equations (62) and (63). Since the fiber is
chemically inert, the degree of the conversion of the matrix is equal to that of the composite.
Eliminating mkmatrix from equations (59) and (62) and substituting the resulting equation and
equation (63) into equation (61), the degree of conversion can be rewritten as

E ~,.~1 c,.(Wo,~ - W . , )
c - 1 - E',~I wf, (64)

S o l i d m a s s loss. Let R F represent the mass fraction of the composite that is resin at virgin as
follows:
matrix
mo
RF = (65)
poV,~

The mass of the degrading matrix can be obtained by inserting equation (63) into equation (61)
and eliminating momatrix from the resulting equation and equation (65):

m matrix= [ c 4 - ( 1 - c ) ~ Wfk]RFpogo. (66)


k=l

We recall that Ms represents the rate of solid mass loss in a unit volume of the mixture. Since
carbon is chemically inert, we have
Fh rnatrix
M~ - ~ (67)

Inserting equation (66) into equation (67) and using equations (20), (52) and (64) with the
small deformation assumption, we obtain

1
M,- - - R F p o Z ~ k ( W o k - Wfk). (68)
1 + ekk

The mass of solid residue at char is m~natrix + m carb°n. The density of the
S h r i n k a g e at char.
composite at char is measured at room temperature and is defined by
m ~ natrix -4- m carb°n
pf = (69)
v~

The mass of carbon, m carb°n, can be obtained by subtracting the original mass of matrix, m omatrix
and the original mass of water, mowater, from the original total mass, poVo. Use of equations (56)
at virgin and (65) leads to
m carb°n (1 - R F - y o ) p o V o .
= (70)

Eliminating mo~atr~x from equations (63) and (65), inserting the resulting equation and equation
Thermomechanical model for chemically decomposing composites 123

(70) into equation (69), and then eliminating V~ from the obtained equation and equation (19)
at char, we obtain

~f= [RF ~ Wfk + (a - R F - 3"o)] P°- I. (71)


k=l Pf

Equation (71) provides the formula for calculating shrinkage at char from available data of the
composite at virgin and at char.
Porosity model. The porosity due to thermochemical decomposition is usually considered as
a linear interpolation of its value at virgin and that at char as described by equation (72):
q~l = Cq~o+ (1 - c)q~f. (72)

State equation of gas


The state equation of gas specifies a relation among three state parameters, i.e. temperature,
pressure, and density. Symbolically the equation is expressed as
/~g =/~g(p, T). (73)
Equation (73) leads to
/~g 1 .
L=-~-ggp - fgJ ", (74)

~,kpg~ = ~gP,k -- fgT, k, (75)

where kg and fg, respectively, represent the bulk modulus and the coefficient of thermal
expansion of a gas defined by

1 _ 1 Ot~ 1 Ot~ (76)


kg-~g Op' fig- ~g OT"
For an ideal gas, the state equation becomes
mw
t~g =R-~P. (77)

As the decomposition process proceeds, different types of gas species are produced and the gas
composition changes with temperature. Therefore, the average molecular weight of the gas mw
is a function of temperature. Inserting equation (77) into equation (76) yields
1 1 dmw
kg = p, fig- T mw dT " (78)

Fourier' s law
Fourier's law shown in equation (79) is used for heat conduction:

qi = - k q T 4, (79)
where kij are the components of effective thermal conductivity tensor of the composite.

GOVERNING EQUATIONS

There are four fundamental variables in this formulation. They are the overall strain eij, pore
pressure p, temperature T, and moisture content 3'- A set of governing equations for these
variables can be obtained by inserting the material response equations into the balance law
equations.
124 Y. WU and N. KATSUBE

M a s s continuity equation
In order to express the mass continuity equation (14) in terms of primary variables, the rate
of change of porosity is needed. This can be obtained by taking the time derivative of equation
(54):

+ ( 6 a - (o,8~j)[eij - 6 - f i a ( T - To) + ( 6 , , - a , , ) G ~ , P l

+ (~,,~ - ~ , ~ i ~ ) [ ( ~ , - o , ~ , ) G ~ , p - ~,,c,~,p]. (80)

The components of elastic compliance tensor C~jk~,as well as the poroelastic coefficients a a, are
generally functions of temperature T. Since temperature is a function of time, Cak, and a~j are
also functions of time. Therefore, the time derivatives of Cak~ and a a are present in the above
equation. Replacing ~blekk by ~bekk under the assumption of small deformation [see equation
(22)], equation (80) can be rewritten as

1
= {4,, + (-,j - ~, ,sij)[~ij - ~'~j - ,eft" + (,sk, - ,~.)c,j.,p]
1 + ekk
+ &ijea - (&ij - (a, 6ij)[G + f l i y ( T - To)]
+ [(6k, + ~, 6 k , - 2ak,)6ijC0k, + (a,j -- qh 3ij)(3k,- Otk,)Cak,- ~bfffa - o~,j)C.,,jIp}. (81)
The velocity gradient of the porous solid skeleton is related to the rate of the overall
deformation by
v(O
k,k ~-- ~kk" (82)

Multiplying equation (14) by (1 + ekk) and inserting equations (58), (68), (74), (81) and (82)
into the resulting equation, the mass continuity equation becomes

a.~'0 + [(1 + e**)~+ (a,j - ~, 6,j)(6.,- a.,)G~.,]p - [(1 + ek.)~,Sg + (aij - ~, 8ij)/3ol'it" + 6'oe,j
k.
-~ [(~kl q- qO1 ~kl -- 20tkl)~ijCijkl + (OCij-- qo, ~ij)(~kl -- ffkl)Oijkl -- (~ij -- ffij)Cnnij~°llP

1 + 8kk
- [6i] - (o,6,j]fla(T - To) + [~gv,],,

- [&a - (~,6ijlKq - (c~,/- qh 6a)~,] + qb, + P o [RFYd,(Wok - Wok)+ 5'] = 0. (83)


Pg

E q u i l i b r i u m equation
Eliminating Lj and ea from equations (47), (48) and (52) and inserting the resulting equation
and equations (58) and (68) into equation (30), under the assumption of small deformation the
equilibrium equation can be expressed as follows:

[Miik,(ek, -- (k, - - / 3 k t A T ) - a,jp]j = - [ R F Y G ( W o k - Wfk) + Y]OoV~. (84)

H e a t transfer equation
Inserting Darcy's law (36), Fourier's law (78) and equations (58) and (68) into the energy
balance equation (46) and ignoring higher order terms, the heat transfer equation can be
rewritten as

[psCps"ff pLCpL'k- qOpgCpg] ~ - ^pg AkiKq


~ ( -~ q~j + p .j ) CpgT,k
= R F p o Z G ( W o k - Wfk)r (¢) + poS/r (~) + (kaT, j),i. (85)
Thermomechanicalmodel for chemicallydecomposingcomposites 125

The bulk density of dry solid and that of water in the equation can be obtained as follows.
From the definition of degree of conversion as in equation (61), the value of solid m a s s m matrix
can be obtained from linear interpolation of its values at virgin and at char. Similarly, by
ignoring the volume change due to deformation and shrinkage, the bulk density of a dry solid
can be approximated as follows:

Ps = c(1 - 3'o)Po + (1 - c)pf, (86)


where (1 - 3'o)po represents the bulk density of the dry solid at virgin. The bulk density of
water is equal to mwater/v. It can be approximated by using equation (56) and the small
deformation assumption as follows:
pL = 3,po. (87)

DISCUSSION

Moisture vaporization
Vaporization of moisture influences not only temperature but also pore pressure and
deformation. On the other hand, vaporization of moisture depends on the thermodynamic
state, which is determined by temperature and pore pressure. In this formulation, this coupled
phenomenon is physically captured thorough the conversations of mass, energy, and linear
momentum. In particular, the rate of moisture vaporization at each point of a continuum is
theroetically determined by the saturation curve of steam and heat available for the phase
change.
McManus [13], McManus and Springer [14] and Kuhlmann [12] included the moisture effect
by using empirical equations where the rate of vaporization is assumed to depend on
temperature or the rate of change of temperature. The constants which appear in these
empirical equations are determined directly from experiments. While these constants include
the effect of pore pressure on moisture vaporization for the particular experiments, no relation
between pore pressure and vaporization is explicitly modeled. Since pore pressure itself
depends on experimental conditions such as sample geometry, moisture content, and heating
conditions, there may be some limitations in predicting moisture vaporization for all possible
conditions. In fact, it has been observed that numerical simulations using these empirical
equations deviate from experimental data when moisture content is high [12]. Our model
captures the physical phenomenon that the thermodynamic state of water is at saturation
during its vaporization process. It differs from these previous works in the sense that the
vaporization rate can be theoretically predicted instead of being specified from empirical data.
However, as described below, the numerical implementation of our model is expected to be
much more complicated than that of previous works.
If there is no vaporization, moisture content remains constant. The fundamental variables are
the overall strain, temperature, and pore pressure. They can be solved based on the coupled
equations of equilibrium, gas mass continuity, and heat transfer. During vaporization, the
thermodynamic state of water is at saturation, and the moisture content changes. Temperature
and pressure are related by the saturation curve of steam. The fundamental variables are the
overall strain, moisture content, and either temperature or pressure. They can be solved based
on the governing equations. In general, initial boundary value problems involve a region with
vaporization, a region without vaporization, and an interface between these two regions.
Therefore, an initial boundary value problem results in a problem of moving vaporization
interfaces.
In our theory, it is assumed that water is initially condensed in the form of a liquid at room
temperature in the porous material. Phase change of the liquid water occurs due to heating,
and the vapor released occupies the pore space. The vaporized water combines with the
126 Y. wu and N. KATSUBE

volatiles created from the thermochemical decomposition and generates pore pressure. The
combined gas escapes from the composite through diffusion. This concept is similar to those of
Kuhlmann [12], McManus [13] and McManus and Springer [14]. Recently, a different approach
has been attempted by Sullivan [10, 11]. In his approach, water is assumed to be contained in a
free volume of the polymer at room temperature. As temperature increases, the mass
concentration of water at any point is determined by a diffusion equation where deformation is
ignored. The average partial pressure of water in the specimen is then calculated from the
average mass concentration of water and the variation in the thermodynamic state of water.
This variation in the thermodynamic state of water is assumed to be approximated by that of
the condensed water at saturation. Sullivan [10,11] has applied this concept to simulate
laboratory tests and reasonable results have been obtained.

Constitutive equations and mechanical properties


From equations (47) and (48), the constitutive relation for the overall response of the porous
composite can be rewritten as

eij = Cijkt'gkt + aktCijktp + flij( T - To). (88)


In the rate form of this constitutive equation, the time derivatives of material properties are
included as in equation (80). In Weiler [16], however, the rate form of the constitutive equation
used is
~,j = Cijk,(tk, + ~k,P) + ]3,jT (89)
where the derivatives of material properties with respect to time are ignored. It is stated [16]
that if C~jkt and a~j depend only on temperature, this equation can be integrated to yield the
total form as in equation (88). This may be true if temperature is independent of time.
However, in many applications temperature is a function of time. Therefore, equation (88) and
equation (89) are not equivalent. Since material properties are measured at fixed temperatures,
the total form of the constitutive relation as in equation (88) should be used. In Sullivan [3] and
Sullivan and Salamon [4, 5], the total form as in (88) is employed for the equilibrium equation.
However, the rate form of equation (89) is implicitly used for the gas mass continuity equation.

Mass balance
In McManus [13] and McManus and Springer [14], the notion of a control volume fixed in
space has been used in deriving the gas mass continuity equation. If a porous solid material
inside the control volume expands, a portion of the gas mass which was originally inside the
control volume will be located outside the control volume. Therefore, in general, deformation
of a porous solid material influences gas mass storage inside the control volume. While
deformation of composites is separately analysed based on the conventional elasticity problem,
this effect is ignored in their formulation. Similarly, in the energy balance equation (43), heat
convection associated with the solid mass flow from a fixed control volume [see (p~v~l)Us +
pLV~l)Ue).k] is neglected. In addition to the above simplification, in solving a one-dimensional
boundary value problem, gas mass storage increase is assumed to be small compared to gas
mass flow. Therefore, gas mass flow is approximated by gas generation, which depends on
temperature or rate of change of temperature. Combining this assumption with the energy
balance, an uncoupled temperature equation is obtained. While this assumption is very useful
in decoupling governing equations, it may not be accurate.
In Kuhlmann [12], the gas volume continuity equation is employed instead of the gas mass
continuity equation. The form of this equation is similar to the one presented in our work if
moisture is ignored. In expressing the continuity equation in terms of fundamental variables,
however, a simplified form of the constitutive equation has been employed in relating the solid
deformation to overall stress and pressure. In our formulation, equation (51) is employed.
Thermomechanical model for chemically decomposing composites 127

In the gas mass balance by Sullivan [3] and Sullivan and Salamon [4], the rate of gas mass
creation per unit volume of a porous solid material is combined with the rate of change of gas
mass storage. T h e y are denoted by Omg/Ot, t which is assumed to consist of four terms. These
four terms are related to pore pressure, temperature, mechanical strain of a porous solid, and
the degree of char. The t e r m related to the degree of char actually represents the rate of gas
mass creation per unit volume. Therefore, the other three terms represent the rate of change of
gas mass storage per unit volume. In describing the rate of change of gas mass storage per unit
volume, the result by Biot and Willis [6] is directly used in his formulation. However, the result
by Biot is obtained by calculating the gas mass expelled from the unit volume for the case
without phase change. In this case, Biot's equation physically represents the gas mass continuity
equation itself. Therefore, in the formulation by Sullivan [3] and Sullivan and Salamon [4], two
types of gas mass continuity equations, i.e. one for the case with phase change and another for
the case without phase change, are used implicitly. In our notation, the rate of change of gas
mass per unit volume of the composite is expressed by
(i)
1 D(/~gVp) /~zVp + pgVp
V Dt V = ~Opg "J¢-~pg ~- ~[~Pg~kk, (90)

which is identical to the first three terms at the left hand side of the gas mass continuity
equation (14) divided by ~3g. In equation (90a), the rate of change of the gas mass storage
depends on the rate of change of gas density and that of the pore volume. In equation (90b),
the rate of change of the pore volume is further decomposed into the overall strain rate and the
rate of change of porosity. In obtaining the gas mass continuity equation, equation (81) is used
for the rate of change of porosity. In equation (81), if we ignore the thermochemical shrinkage
and further delete the terms associated with the rate of change of material properties (ti 0 and
OOkt), the rate of change of pore volume due to mass removal (~1), and the rate of pore
volume increase due to thermal expansion (~lflkkT"), the expression for the rate of change of
gas mass storage in equation (90) is then reduced to the one which appears in Sullivan and
Salamon [4]. If we further neglect the moisture t e r m in equation (14), the gas mass continuity
equation used by Sullivan and Salamon [4] is recovered.

Acknowledgements--This work has been supported by Nissan Motor Co. Ltd., Aerospace Division. The authors are
grateful for the valuable suggestions and corrections made by reviewers.

REFERENCES

[1] Henderson, J. B. and Wiecek, T. E., J. Composite Mater., 1987, 21, 373.
[2] Buch, J. D., Thermal expansion behavior of a thermally degrading organic matrix composite, The Aerospace
Corporation. Prepared for Space and Missile Systems Organization, Air Force Systems Command, Los Angeles,
1971.
[3] Sullivan, R. M., A finite element model for thermo-chemically decomposing polymers. Ph.D. dissertation, The
Pennsylvania State University, 1990.
[4] Sullivan, R. M. and Salamon, N. J., Int. J. Engng Sci., 1992, 30, 431.
[5] Sullivan, R. M. and Salamon, N. J., Int. J. Engng Sci., 1992, 30, 939.
[6] Biot, M. A. and Willis, D. G., AMSE J. Appl. Mech., 1957, 24, 594.
[7] Kansa, E. J., Perlee, H. E. and Chaiken, R. F., Combust. Flame, 1977, 29, 311.
[8] Sullivan, R. M., J. Composite Mater., 1993, 27, 408.
[9] Salamon, N. J. and Lee, S., Axisymmetric finite element analysis of decomposing polymeric composites and
structures. Proceedings of the JANNAF Rocket Nozzle Technology Subcommittee Meeting, 1994.
[10] Sullivan, R. M., The effect of water on the thermal expansion behavior of FM5055 carbon phenolic. Proceedings of
the JANNAF Rocket Nozzle Technology Subcommittee Meeting, 1994.
[11] Sullivan, R. M., ASMEJ. AppL Mech., 1996, 63, 173.

t In Sullivan [3] and Sullivan and Salamon [4], O/Ot indicates the material time differentiation of a porous solid
(1)
material, which is denoted by D/Dt in this paper.
128 Y. WU and N. KATSUBE

[12] Kuhlmann, T. L., Thermo-chemical-structural analysis of carbon-phenolic composites with pore pressure and
pyrolysis effects. Ph.D. dissertation, University of California, Davis, 1990.
[13] McManus, H. L. N., High temperature thermo-mechanical behavior of carbon-phenolic and carbon-carbon
composites. Ph.D. dissertation, Stanford University, 1990.
[14] McManus, H. L. N. and Springer, G. S., J. Composite Mater., 1992, 26, 206.
[15] McManus, H. L. N. and Springer, G. S., J. Composite Mater., 1992, 26, 230.
[16] Weiler, F. C., Computational Mechanics of Porous Materials and Their Thermal Decomposition, Vol. 136.
ASMEAMD, 1992, p. 1.
[17] Wu, Y. and Katsube, N., Computational Mechanics of Porous Materials and Their Thermal Decomposition, Vol.
136. ASMEAMD, 1992, p. 103.
[18] Wu, Y. and Katsube, N., Mechanics of Materials, 1996, 22, 189.
[19] Green, A. C. and Naghdi, P. M., Int. J. Engng Sci., 1965, 3, 231.
[20] Katsube, N. and Carroll, M. M., ASME J. Appl. Mech., 1987, 54, 35.
[21] Crochet, M. J. and Naghdi, P. M., Int. J. Engng Sci., 1966, 4, 383.
[22] Carroll, M. M., J. Geophys. Res., 1979, 84, 7510.
[23] Katsube, N., Int. J. Solids Struct., 1988, 24, 375.

(Received and accepted 13 May 1996)

You might also like