Telechargé par ounaies mohamed

1-s2.0-S0168874X10000685-main

publicité
Finite Elements in Analysis and Design 46 (2010) 783–791
Contents lists available at ScienceDirect
Finite Elements in Analysis and Design
journal homepage: www.elsevier.com/locate/finel
A finite element model for hygro-thermo-mechanical analysis of masonry
walls with FRP reinforcement
Mehran Khoshbakht , Mark W. Lin
University of Alabama in Huntsville, Department of Mechanical and Aerospace Engineering, Huntsville, AL 35899, USA
a r t i c l e in fo
abstract
Article history:
Received 24 May 2007
Received in revised form
8 October 2009
Accepted 24 April 2010
Modeling the effects of humidity and temperature gradients on the structural behavior of masonry
walls reinforced with fiber reinforced polymer (FRP) composite is of great importance. Study of
interfacial stresses, in particular, is a key factor in predicting the durability of the bond between the
reinforcing FRP laminate and the host masonry. In this paper, a finite element modeling procedure for
analyzing the hygro-thermo-mechanical response of multi-layered structures constructed with
distinctive permeable materials was developed by incorporating structural stress analysis into the
coupled moisture/temperature finite element model based on the governing equations proposed by
Phillips and De Vries. The hygro-thermo-mechanical finite element model was used to analyze the
response of a concrete block reinforced with a unidirectional glass/epoxy FRP composite laminate. The
results demonstrated that the effect of temperature gradient on the moisture distribution, i.e., heatinduced moisture movement, resulted in an accumulation of moisture at the interface, and induced
interfacial stresses even in the absence of a moisture gradient. The presented finite element modeling
procedure can be used to aid with the design of FRP/masonry structure or other similar structures for
minimizing interfacial stresses induced due to the mismatch of moisture swelling and thermal
expansion properties of the constituent materials.
& 2010 Elsevier B.V. All rights reserved.
Keywords:
Moisture and temperature
Structural analysis
FRP reinforced masonry
Multi-layered permeable structure
1. Introduction
Many of aging buildings were constructed using masonry
structures. Most of the existing masonry structures, except a few,
were not reinforced to meet modern engineering codes. Masonry
structures without any reinforcement have inadequate lateral
strength due to the presence of a matrix of adhesive joints of
mortar. Structures constructed with un-reinforced masonry,
therefore, are prone to failure when subject to lateral loading
such as high winds and earthquake events. Among the contemporary methods for strengthening un-reinforced masonry
structures, using fiber reinforced polymer (FRP) composites has
been shown to be an effective and efficient technique [1,2]. Fiber
reinforced polymer composites offer a combination of outstanding properties that are very suitable for masonry wall upgrade
applications. They have low specific weight and excellent
mechanical stiffness and strength. They are also immune to
corrosion and have the ability to be formed in different shapes
and very long lengths.
Though many studies have demonstrated the effectiveness of
the FRP reinforcement for masonry structures [3,4], one important
Corresponding author. Tel.: + 1 310 658 3301.
E-mail address: [email protected] (M. Khoshbakht).
0168-874X/$ - see front matter & 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.finel.2010.04.002
issue that hinders its broad application is the understanding of
environmental temperature and moisture effects on the FRP
reinforcement and the FRP/masonry interface, in particular. The
day-to-day climate changes that these structures have to sustain
for several decades can result in substantial damaging effects on
the structural integrity, especially at the interface between the
masonry substrate and the FRP. The issue of thermo-mechanical
behavior of the building structures is well known. However, the
effect of moisture swelling mismatches between the masonry and
FRP materials, which can induce considerably high level of
interfacial stresses, has not been thoroughly investigated. This
effect may result in disintegration of the FRP/masonry interfacial
bond and damages in both masonry and FRP constituents [5,6].
Since the early 1990s, research on various aspects of using FRP
composites as upgrade materials on un-reinforced masonry
structures has received much attention [7–12]. To determine
the effect of the environment on the mechanical behavior of such
structures, it is necessary to derive a mathematical model for the
moisture diffusion and heat transfer phenomenon in multilayered porous structures. The most accepted governing equations capable of describing hygro-thermal problem for a porous
solid were derived by Philip and De Vries [13] and Luikuv [14].
Their mathematical models include two coupled partial differential equations for moisture content and temperature dependent
variables, and are applicable to a single-material structure.
784
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
These equations have been widely used as the basis for many
studies in this area with various simplifications according to the
specific physics and geometry of the problem investigated. Some
studies have been done to find the analytical solutions for the
coupled moisture/temperature problem in porous media with
regular geometries [15–17]. In all these studies, the material
transport properties were considered constant in order to obtain
closed-form analytical solutions, even for structures with simple
geometries. Other researchers have alternatively used various
numerical techniques to solve the coupled hygrothermal problem
[18–20]. Several published works are also available regarding the
analysis of the mechanical behavior of the porous structures
caused by the moisture variations through the medium [21–24].
These studies, however, have limitations such as specific simple
geometry, constant material properties, monolithic material, nonpermeability of part of the structure, or one-dimensional moisture
migration analysis.
In spite of these valuable studies, to date, there is little
research that addresses the three-dimensional hygro-thermomechanical problem of a multi-layered structure constructed of
distinctive permeable materials with general moisture/temperature dependent material properties. Due to the complexity of the
governing equation for the hygro-thermo-mechanical problem as
well as the geometry of these structures, obtaining a closed-form
analytical solution is very difficult, if possible. Simplification of
the equations and the geometry would result in unrealistic
solutions with little or no practical value. On the other hand,
experimental study of the long-term environmental effects on
such structures would be very time consuming and expensive.
Using finite element technique to carry out comprehensive
studies for the long-term durability problem of FRP composites
reinforced masonry structures is a practical alternative.
In this paper, a finite element model, previously presented by
the authors to analyze the hygrothermal behavior of an FPR/
concrete wall structure [25] is augmented to develop a comprehensive model for hygro-thermo-mechanical analysis of this type
of structures. The structural analysis formulation is incorporated
into the hygrothermal model to determine the effects of moisture
swelling and thermal expansion on the mechanical response of
the wall structure. The developed hygro-thermo-mechanical
model is then implemented to analyze the behavior of a typical
concrete slab with FRP laminate reinforcement under various
temperature and humidity gradient loads. The phenomenon of
heat-induced moisture movement is particularly investigated to
evaluate its resulting interfacial stresses at the FRP/concrete
bonding surface.
2. Finite element implementation of the hygrothermal
problem
rc
@T
¼ r ðlrTÞ þ hv r ðDyvap ryÞ,
@t
rc
ð1Þ
ð2Þ
where y is the moisture content (kg/m3), T is the temperature (K),
Dy is the moisture diffusion coefficient (m2/s), DT is the thermal
moisture diffusion coefficient (kg/(m s K)), rc is the volumetric
heat capacity (J/(m3 K)), l is the thermal conductivity (W/(m K)),
hv is the latent heat of vaporization (J/kg), Dyvap is the isothermal
vapor diffusivity (m2/s).
@T
dP
¼ r ðlrTÞ þ hv r rðhPs Þ ,
@t
m
ð4Þ
where h is the relative humidity of the surrounding air
(dimensionless), m is the vapor resistivity of the porous medium
(dimensionless), dP is the vapor permeability of the air (kg/
(m s Pa)), Ps ¼ P0 exp½ðDH=RÞðð1=TÞð1=373:15ÞÞ is the saturation
vapor pressure (Pa), in which P0 ¼101,325 (Pa) is the pressure of a
standard atmosphere, and DH/R ¼5201 (K) measured for liquid
water.
For a multi-layered structure, since the moisture content is
discontinuous at the interface of different layers of constituent
materials, a change of variable approach was taken by Lin et al.
[27] to substitute moisture content, y, with relative humidity, h,
which resulted in the following form for the moisture diffusion
equation:
Ch
@h
¼ r ðD h rhÞ þ r ðD T rTÞ,
@t
ð5Þ
where
Ch ¼
@y
,
@h
ð6Þ
Dh ¼
@y
dP
D þ Ps ,
@h h m
ð7Þ
DT ¼
dp DH
m RT 2
hPs ,
ð8Þ
in which Dh is the moisture diffusion coefficient of the material
expressed in terms of relative humidity using sorption isotherm
relation.
The second term in the heat transfer equation, Eq. (4), which is
associated with latent heat was shown to be insignificant for the
temperature range considered and, therefore, was omitted. As the
result, the change of variable did not affect this equation. Thus,
the heat equation yields a simple form as
rc
The coupled moisture migration and heat transport problem in
porous media was formulated by Philip and De Vries [13] in the
form of the following system of differential equations:
@y
¼ r ðDy ryÞ þ r ðDT rTÞ,
@t
The transport coefficients are to be determined experimentally. Among these coefficients, there is a lack of vigorous
definition for the coefficient of thermal moisture diffusion, DT.
Therefore, American Society for Testing and Materials (ASTM) in
2001 suggested a modified version for this system of equations as
the standard tool for moisture analysis in building envelopes [26].
In the proposed mathematical model, the governing equations
take the modified form as
@y
dP
¼ r ðDy ryÞ þ r ð3Þ
rðhPs Þ ,
@t
m
@T
¼ r:ðlrTÞ,
@t
ð9Þ
The values or functional forms of the various coefficients in the
governing equations as well as sorption isotherm relation for
typical concrete and FRP materials have been experimentally
characterized and collectively summarized [28]. Applying the
Galerkin finite element formulation to the coupled humidity and
heat transport equations yields the following finite element
equation in terms of nodal humidty/temperature vector
( )
hi
ði ¼ 1,2, . . . , rÞ, with r being the total number of nodes
Ti
per element,
9
8
@hi >
>
( )
>
>
=
<
hi
@t
þ½D e ¼ fQe g,
½Ce @T
>
>
Ti
i
>
>
;
:
@t
ð10Þ
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
where ½D e ¼
R
T
½D½B dV is the element humidity and heat
2
3
D hx
0
0
D Tx
0
0
6
7
D hy
0
0
D Ty
0 7
6 0
6
7
6 0
D hz
0
0
D Tz 7
0
6
7
diffusion matrix, ½D ¼ 6
7 is the
6 0
0
0
lx
0
0 7
7
6
6 0
ly
0 7
0
0
0
4
5
lz
0
0
0
0
0
V ½B
material humidity and heat diffusion matrix, [B]¼LNT, in which
(
)T
N1 0 N2 0 Nr 0
is the element
fNðx,y,zÞg ¼
0 N1 0 N2 0 Nr
shape function,
@
@
@
@
@
@
the
differential
operator,
@x @y @z @x @y @z
" #
0
R Ch
fNgfNgT dV is the element humidity and heat
½Ce ¼ V
0 rc
R
capacity matrix, fQe g ¼ S fNgq dS is the element boundary
humidity and heat flux, qn is the boundary humidity/heat flux.
fLg ¼
3. Hygro-thermo-mechanical finite element model
To date, none of the commercial finite element software
packages has an available capability to model coupled structural/
moisture diffusion/heat transfer problem in a porous medium.
Therefore, in order to model the hygro-thermo-mechanical
problem using finite element technique, moisture-induced strain
and thermal strain need to be incorporated into the stress–strain
constitutive relation of the material in the finite element code.
Since hygro-thermal induced strains of most masonry materials
and FRP are relatively small and exhibit no permanent deformations, the linear elastic stress–strain relation is applicable. It is
also noted that stress-induced effects on thermal conductivity and
moisture diffusivity of the materials are not taken into consideration in the current study, since such effects appear to be negligible
for the stress levels considered.
The linear elastic constitutive equation including the moisture
swelling and thermal strains can be expressed as
fsg ¼ ½Dfeelastic g ¼ ½Dfetotal eswelling ethermal g,
ð11Þ
where fsg ¼ fsx sy sz txy tyz txz gT is the stress vector, [D] is the
stiffness matrix of the material, etotal is the total apparent strain,
feelastic g ¼ fex ey ez gxy gyz gxz gTelastic is the elastic strain vector,
eswelling is the moisture-induced strain, and ethermal is the thermal
strain. For an orthotropic material, such as a unidirectional FRP
composite laminate, the moisture and thermal induced strain
vectors have the form
feswelling g ¼
n
b1 ðyyref Þ b2 ðyyref Þ b3 ðyyref Þ 0 0 0
oT
,
ð12Þ
fethermal g ¼
n
a1 ðTTref Þ a2 ðTTref Þ a3 ðTTref Þ 0 0 0
oT
,
ð13Þ
where bi, i¼ 1, 2, 3 are moisture swelling coefficients in the
principle material directions, ai, i¼1, 2, 3 are the thermal
expansions coefficients in the principle material directions, and
yref and Tref are the reference moisture content and the reference
temperature, at which the moisture and thermal induced strains
are zero, respectively. Apparently, bi has the same magnitude in
all directions for an isotropic material, such as masonry. Likewise
785
ai also has the same magnitude in all directions for isotropic
masonry materials.
The moisture swelling coefficient for a porous medium is
generally measured in terms of moisture content. Since the
moisture transport problem in a layered structure was converted
to a problem in terms of humidity potential, this coefficient also
needs to be modified accordingly. Due to the fact that the
moisture-induced strain calculated from the moisture content
state variable and from the humidity potential state variable must
yield the same induced strain, the following relation must hold:
eswelling ¼ b y ¼ b h,
ð14Þ
where bn is defined as the humidity swelling coefficient.
It is noted that in the case where b is given in terms of specific
moisture content, the humidity swelling coefficient, bn, has the
form
y
b,
b ¼
ð15Þ
rh
where r is the mass density of the material.
In the present study, the swelling coefficients, b, for both
concrete and FRP is treated as a constant, since only linear
response range is considered. The resulting humidity swelling
coefficient, bn, is dimensionless and represents strain per unit
relative humidity. It should be noted that because the moisture
content, y, is generally a nonlinear function of the relative
humidity, h, based on the characteristic sorption isotherm
property, the resulting swelling coefficient bn becomes a nonlinear function of, h.
The total strain vector can be related to displacements by the
displacement–strain relation as
8 9
>
<u>
=
T
fetotal g ¼ fex ey ez gxy gyz gxzz g ¼ fL g v ,
ð16Þ
>
: >
;
w
where Ln is a
8
@
>
>
>
> @x
>
>
>
<
0
L ¼
>
>
>
>
>
>
>
:0
differential operator defined as
9
@
@ >T
>
0 0
0
>
@y
@z >
>
>
>
=
@
@
@
0
0
:
@y
@x @z
>
>
>
>
@
@
@ >
>
>
0
0
;
@z
@y @x
ð17Þ
Therefore, the finite element equation for the structural
analysis incorporating the moisture and temperature strains
results in the following form:
½KfUg ¼ fFB g þ fFS g þfFthermal g þ fFswelling g,
ð18Þ
T
where fUg ¼ fu1 v1 w1 u2 v2 w2 . . . ur vr wr g is the nodal
R
displacement vector, ½K ¼ V ½B T ½D½B dV is the element stiffR
T
ness matrix, fFB g ¼ V ½N fFb g dV is the element body force vector,
R
FS ¼ S ½N T fFs g dS is the element force vector due to the surface
R
loading, fFthermal g ¼ V ½B T ½Dfethermal g dV is the element force
R
vector due to the thermal expansion, fFswelling g ¼ V ½B T
½Dfeswelling g dV is the element force vector due to the moisture
swelling.
T
In the above equations ½B ¼ fLgfN g , in which Ln was defined
in Eq. (17) and
9T
8
0 N2 0
0 : : : Nr 0
0 >
>
=
< N1 0
0 N2 0 : : : 0 Nr 0
0 N1 0
fN ðx,y,zÞg ¼
,
>
>
: 0
0 N2 : : : 0
0 Nr ;
0 N1 0
is the shape function matrix. To implement the coupled hygrothermo-mechanical finite element analysis, the FEMLABs commercial finite element analysis package, which allows its user to
786
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
modify the predefined finite element formulation, was employed.
Considering the highly nonlinearly coupled nature of the problem,
the damped Newton method provided in the nonlinear solver of
FEMLAB software was used [29]. The software approximates the
nonlinear model with a linear model using initial values of
humidity, temperature, and displacement and calculates the
coefficients for the linearized system. The linear direct solver
(UMFPACK) was selected to solve the linearzied system. The
solution from the current linearized system is used to re-calculate
the coefficients and establish an updated system of linearized
equations. The process is repeated until a converged solution is
obtained. The linear direct solver uses the permutation of the
rows of the system of equations to achieve a more stable and
more accurate solution.
To solve the transient problem, the FEM discretization of the
time-dependent PDE problem, was carried out using the ‘‘method
of lines’’ technique. To solve the resulting system, DASPK solver
[30] was used with the default values of 0.01 and 0.001 for
relative and absolute tolerances, respectively.
Units: meter
0.2
0.15
0.1
0.05
0
0.4
0.35
-0.05
0.2
0.3
0.25
0.15
Z
0.2
0.1
0.15
0.1
0.05
Y
X
0
0.05
0
Fig. 2. Schematic of the physical model with displacement boundary conditions.
4. Case study of a concrete slab with FRP laminate
reinforcement
A three-dimensional model of a concrete slab with dimensions
of 396 194 194 mm3 partially covered with an FRP reinforcement laminate was considered as a basis to carry out the coupled
hygro-thermo-mechanical analysis. The model configuration
along with the assigned coordinate system is shown in Fig. 1.
The thickness of the FRP laminate was assigned as 19.4 mm,
which corresponds to 10 layers of a typical woven glass/epoxy
composite lamina and is about 10% of the thickness of a typical
concrete masonry unit block used in building construction. The
length of FRP in X-direction was chosen as 60% of the length of the
concrete block, which is equal to 238 mm. Since applying FRP
laminates on the inner surface of a wall structure is a common
practice, outdoor and indoor boundary conditions were applied
on the concrete and FRP sides, respectively. The humidity and
temperature boundary conditions were considered as hin ¼50%
and Tin ¼20 1C at FRP side, and hout ¼97% and Tout ¼10, 20, 40, 50,
and 60 1C at the concrete side. The lateral, top, and bottom
surfaces of the concrete block (normal to Z and X coordinates,
respectively) and the top and bottom faces of the FRP laminate
were considered as insulated. Under the consideration of such
boundary conditions, the moisture and heat transport problem in
the model became essentially two-dimensional, whereas the
structural analysis was still a three-dimensional analysis.
The schematic of the model with prescribed displacement
constraint boundary conditions is shown in Fig. 2. The nodal
194 mm
396 mm
238 mm
194 mm
Z
Y
X
Fig. 1. Three-dimensional model of the FRP reinforced concrete block.
displacement uz on all nodes on the Z¼ 0 surface is constrained
based on the consideration of symmetry. The displacements in the
X-direction for the nodes on the central edges of the model with
Y¼ 194 and 194 mm are also restrained. Finally, to prevent rigid
body motion in the Y-direction, Y-displacement of the node at the
coordinates of X¼198 mm, Y¼194 mm, and Z¼ 0 is also
restrained. The values or functional forms of the hygro-thermomechanical material properties adopted in the current analysis
are given in the Appendix.
5. Results and discussion
The resulting steady-state relative humidity and corresponding moisture distributions along the centerline of the block in the
FRP and concrete, i.e., X¼198 mm, obtained for various outdoor
temperatures are shown in Figs. 3 and 4, respectively.
It is observed that increasing the outdoor temperature causes
the overall profile of the relative humidity/moisture content to
shift upward when the outdoor temperature is greater than the
indoor temperature, and to shift downward otherwise. This
behavior can be explained considering the temperature distribution shown in Fig. 5. An apparent abrupt change can be seen in the
slopes of the temperature distribution curves at the FRP/concrete
interface. For the cases where the outdoor temperature is greater
than the indoor temperature, the slope increases. The positive
second derivative of the temperature profile in these cases
behaves as a moisture source in the second term of Eq. (5),
which drives the relative humidity/moisture content to a higher
value. On the other hand, the negative second derivative of the
temperature profile, as in the case where the outdoor temperature
is lower than the indoor temperature, behaves as a moisture sink,
which drives the relative humidity/moisture content to a lower
value. This phenomenon is known as ‘‘heat-induced moisture
movement’’.
Transient analyses for various thermal gradient conditions
were also carried out. The resulting relative humidity distributions at various time instants along the centerline of the block for
the case with Tin ¼ 20 1C and Tout ¼40 1C are shown in Fig. 6. Again,
humidity boundary conditions were prescribed as hin ¼50% and
hout ¼97%.
At all time instants, apparent abrupt changes in the slopes of
the temperature distribution profile at the FRP/concrete interface
are present due to the distinctive coefficients of thermal
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
787
Y (mm)
-194
-155.2
-116.4
-77.6
-38.8
0
38.8
77.6
116.4
155.2
194
100
Relative Humidity (%)
90
T=60 C
T=50 C
80
T=40 C
T=30 C
T=20 C
70
FRP
Concrete
60
T=10 C
50
40
0
3.88
7.76
11.64
15.52
19.4
Y (mm)
Fig. 3. Steady-state humidity distribution along the centerline (X ¼ 198 mm) for various outdoor temperatures.
Y (mm)
-77.6
-38.8
0
38.8
77.6
116.4
155.2
194
12
120
11
110
10
100
9
90
8
80
T=60 C
7
T=50 C
6
T=40 C
5
T=30 C
4
T=20 C
3
FRP
Concrete
T=10 C
2
1
70
60
50
40
30
20
10
0
0
0
3.88
7.76
11.64
15.52
19.4
23.28
27.16
31.04
34.92
3
-155.2 -116.4
Mositure Content in Concrete (kg/m )
Moisture Content for FRP (kg/m3)
-194
38.8
Y (mm)
Fig. 4. Steady-state moisture content distribution along the centerline (X ¼198 mm) for various outdoor temperatures.
Y (mm)
-194
-155.2 -116.4
-77.6
-38.8
0
38.8
77.6
116.4
155.2
194
70
T=60 C
Temperature (°C)
60
T=50 C
50
T=40 C
T=30 C
40
FRP Concrete
T=20 C
30
T=10 C
20
10
0
0
3.88
7.76 11.64 15.52 19.4 23.28 27.16 31.04 34.92 38.8
Y (mm)
Fig. 5. Steady-state temperature distributions along the centerline (X ¼198 mm) for various outdoor temperatures.
conductivity for the FRP and concrete materials. This trend is
similar to the temperature distribution profile shown in Fig. 5 for
the steady-state condition. As a result, the aforementioned effect
of thermal moisture flux on the moisture movement distribution
is also evident at all the time instants shown. Therefore, for the
case shown in Fig. 6, where the outdoor temperature is greater
than the indoor temperature, the positive second derivative of the
temperature profile has functioned as a moisture source, which
has driven the relative humidity/moisture content to higher levels
for all the curves. Since the major change in local temperature
788
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
Y (mm)
Relative Humidity (%)
-194 -155.2 -116.4 -77.6 -38.8
100
0
38.8
77.6 116.4 155.2
194
100
Steady
State
10 months
90
90
80
80
70
70
60
60
1 month
50
5 days
5 months
3 months
50
FRP
40
0
3.88
Concrete
2 months
1 day
40
7.76 11.64 15.52 19.4 23.28 27.16 31.04 34.92 38.8
Y (mm)
Fig. 6. Transient relative humidity distributions along the centerline (X¼ 198 mm) for Tin ¼20 1C and Tout ¼40 1C.
600
500
Tout = 60 °C
τzy (kPa)
400
Tout = 40 °C
300
200
Tout = 20 °C
100
Tout = 10 °C
0
-100
40
60
80
100
120
140
160
180
200
Z (mm)
Fig. 7. Interfacial shear stress, tyz distribution along the z-axis at X¼ 198 mm on the FRP laminate.
gradient occurs at the interface, while this quantity has very small
magnitudes at the locations away from the interface, a drastic
increase in the relative humidity resulting from the change in
temperature gradient appears only at the interface.
The hygro-thermal induced interfacial shear stress tyz on the
FRP laminate along the Z-axis for various outdoor temperatures is
presented in Fig. 7 for the humidity/temperature loading obtained
from the steady-state condition. It can be seen that the level of
this stress, generally, increases with the increasing temperature
gradient. Noting that the humidity gradient is the same for all
these cases, it is apparent that this increase is caused by two
factors. The first factor is the increase in the thermal expansion
due to higher temperature gradients. The second factor is the
increase in the moisture swelling due to heat-induced moisture
movement.
It order to quantify this effect, two additional analyses were
conducted. In the first analysis, no moisture gradient was
prescribed, i.e., hin ¼hout ¼50%, while a temperature gradient
was given as Tout ¼60 1C and Tin ¼20 1C. In the second analysis,
the temperature and humidity boundary conditions were kept the
same, but the thermal expansion effect was eliminated by
assigning zero thermal expansion coefficients for concrete and
FRP. It is apparent that in the second analysis the only induced
strain is due to the heat-induced moisture movement.
Fig. 8 shows the humidity distribution for the second analysis
along the centerline of the block, where X ¼198 mm and Y
changes from 19.4 to 194 mm. It is clearly shown that even in
the absence of the humidity gradient; there exists an increase in
the humidity close to the interface of the concrete and FRP
laminate, which is caused by the heat-induced moisture
movement.
The comparison of the shear stress tyz profile for these three
analyses is shown in Fig. 9. It reveals that even with no moisture
gradient and thermal expansion effect, there is still some level of
the interfacial shear stress caused by heat-induced moisture
movement. Noting the maximum value of this stress, it can be
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
789
90
Relative Humidity (%)
80
70
60
50
40
30
Concrete
FRP
20
10
0
-20
0
20
40
60
80
100
120
140
160
180
200
Y (mm)
Fig. 8. Relative humidity distribution at the centerline for the case with no relative humidity gradient.
600
500
Relative humidity gradient (97% / 50%)
With thermal expansion effect
τzy (kPa)
400
300
No relative humidity gradient
With thermal expansion effect
200
100
No relative humidity gradient
Without thermal expansion effect
(Only heat-induced moisture movement)
0
-100
40
60
80
100
120
140
160
180
200
Z (mm)
Fig. 9. Comparison of the interfacial shear stress, tyz distributions along the z-axis at X ¼ 198 mm on the FRP laminate.
2000
τxy (kPa)
1500
1000
Tout = 60 °C
Tout = 40 °C
500
Tout = 20 °C
Tout = 10°C
0
-500
198
208
218
228
238
248
258
268
278
288
298
308
318
X (mm)
Fig. 10. Interfacial shear stress, txy distribution along the x-axis at Z ¼194 mm on the FRP laminate.
concluded that the stress due to this phenomenon comprises
about 25% of the total induced stress.
The profile of the interfacial shear stress, txy, for various
outdoor temperatures along the x-axis at Z¼194 mm is shown in
Fig. 10. Similar trend to that of tyz is also observed fortxy, although
the stress level is less severe.
The distribution profile of the interfacial peeling stress, sy, for
various outdoor temperatures along the x-axis at Z ¼194 mm in
790
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
Sorption isotherm relation
1200
y ¼ 226:68h3 247:75h2 þ 123:45h þ:1076 ðkg=m3 Þ:
1000
ðA:1Þ
Moisture diffusion coefficient
σy (kPa)
800
Dy ¼ 10:01942y10:5 ½m2 =s,
600
Tout = 60 °C
400
Tout = 40 °C
kg
for 0 o y
o 200:
m3
ðA:2Þ
Thermal moisture diffusion coefficient
200
Tout = 20 °C
5
4
3
DT ¼ r0 ð1:67 108 y 3:99 106 y þ 2:58 104 y
4:14 10
3 2
y þ:216y:035Þ10
13
ðkg=ðm s KÞÞ,
ðA:3Þ
0
-200
Tout = 10 °C
in which r ¼2,200 (kg/m3).
Volumetric heat capacity
rc ¼ r0 c0 þ 4187y ðJ=ðm3 KÞÞ,
-400
198 208 218 228 238 248 258 268 278 288 298 308 318
X (mm)
Fig. 11. Peeling stress, sy distribution along the x-axis at Z ¼194 mm on the FRP
laminate.
where c0 ¼940 (J/(kg K)) for dry material.
Thermal conductivity
l ¼ 2:74 þ :0032y ðWatt=ðm KÞÞ:
6. Conclusion
A finite element modeling procedure for analyzing the hygrothermo-mechanical behavior of multi-layered structures consisted of distinct permeable materials was developed by incorporating the thermal and swelling strains in the coupled
moisture/temperature finite element model based on the governing equations presented by Philip and De Vries.
The developed model was then used to analyze the structural
response of a concrete block reinforced with a unidirectional
glass/epoxy FRP composite laminate partially covering one lateral
surface of the block. The reinforced concrete block was subjected
to humidity and temperature gradient loads resulted from the
difference of indoor and outdoor environmental conditions. The
results showed that significant levels of interfacial shear stresses
were induced by three factors: temperature gradient, moisture
gradient and heat-induced moisture movement. The phenomenon
of the heat induced moisture movement was clearly observed in
the form of moisture accumulation at the interface of the concrete
and FRP, even in the absence of a moisture gradient load. This
moisture accumulation can cause up to 25% of the total induced
stress under the general loading of temperature and humidity
gradients. The effect of heat-induced moisture movement on
interfacial stresses of multi-layered structures was shown in the
presented study by incorporating the temperature gradient term
in the finite element formulation of the moisture diffusion
equation, which could not be detected otherwise.
The developed finite element modeling procedure provides a
general method for hygro-thermo-mechanical analysis of multilayered permeable structures. It can be used as an aid with the design
of FRP/masonry structure or other similar structures for minimizing
interfacial stresses due to the mismatch of moisture swelling and
thermal expansion properties of the constituent materials.
The functional forms or values of material parameters for
concrete are
ðA:6Þ
Thermal expansion coefficient
a ¼ 9 106 ð1=KÞ
ðA:7Þ
The functional forms or values of material parameters for FRP
laminate are
Sorption isotherm relation
y ¼ 7:53h4:3 ðkg=m3 Þ:
ðA:8Þ
Moisture diffusion coefficient
DH
1013 ½m2 =s,
Dy ¼ 2659:7ð1:865y þ 49:125Þ Exp RT
kg
o 10,
ðA:9Þ
for 0 o y
m3
where DH/R ¼2626.03 (K).
Thermal moisture diffusion coefficient
dp h
P ðkg=ðm s KÞ,
DT ¼
m T s
ðA:10Þ
where m ¼vapor resistivity¼40,000,
0:81
T
dP ¼ 1:83 1010 ðkg=ðm s PaÞÞ,
273:15
DH 1
1
ðPaÞ,
Ps ¼ P0 exp R T 373:15
in which P0 ¼101,325 (Pa) is the pressure of a standard atmosphere, and DH/R¼5201(K) measured for liquid water.
Volumetric heat capacity
rc ¼ nf rx cx þ ð1nf Þre ce ðJ=ðm3 KÞÞ,
ðA:11Þ
where rxcx is the dry volumetric heat capacity of the
fiber¼ 0.209 107 (J/(m3 1C)) for glass fiber, rece is the volumetric
heat capacity of the epoxy matrix¼ r0c0 + 4187y, in which
r0c0 ¼0.36 107 (J/(m3 1C)) for dry material, nf is the fiber volume
fraction E0.6.
Thermal conductivity
Longitudinal direction
lL ¼ nf lf þð1nf ÞlT ðW=ðm KÞÞ,
Appendix
ðA:5Þ
Moisture swelling coefficient
b ¼ 0:00762:
the FRP laminate is shown in Fig. 11. It can be seen that the
maximum peeling stress occurs at the free edge, which is the
resultant of the warping deformations of the concrete and FRP
laminate about both x- and z-axis, as expected.
ðA:4Þ
Transverse direction
1 þ Znf
lTrans ¼
lT ,
1Znf
ðA:12Þ
ðA:13Þ
M. Khoshbakht, M.W. Lin / Finite Elements in Analysis and Design 46 (2010) 783–791
where lf ¼1.05 (W/(m K)) is the thermal conductivity of a glass
fiber, lT ¼ 9:522 1011 T 4 þ 1:111 107 bT 3 4:057 105 T 2
þ4:263 103 T þ :382 is the thermal conductivity of the epoxy
matrix, Z ¼ ððlf =lT Þ1Þ=ððlf =lT Þ þ1Þ.
Moisture swelling coefficient
b1 ¼ 0, b2 ¼ b3 ¼ 0:6 106 :
ðA:14Þ
Thermal expansion coefficient
a1 ¼ 6 106 ð1=KÞ, a2 ¼ a3 ¼ 21 106 ð1=KÞ
ðA:15Þ
References
[1] H. Saadatmanesh, Extended service life of concrete and masonry structure
with composites, Construction and Building Materials 11 (No. 5–6) (1997)
327–335.
[2] T.C. Triantafillou, Composites: a new possibility for the shear strengthening of
concrete, masonry, and wood, Composites Science and Technology 58 (1998)
1285–1295.
[3] O.S. Marshall, S.C. Sweeney, and J.C. Trovillion, Seismic rehabilitation of
unreinforced masonry walls, in: Proceedings of the Fourth International
Conference on Fiber Reinforced Plastics for Reinforced Concrete Structures
(FRPRCS-4), October/November 1999, Baltimore, MD, pp. 287–295.
[4] G.K. Al-Chaar, H.A. Hassan, Dynamic response and seismic testing of CMU
walls rehabilitated with composite material applied to only one side,
Structures & Buildings Journal, Special Issue: Dynamic Behavior and Earthquake Design 152 (No. 2) (2002) 135–146.
[5] C. Au, Moisture degradation in FRP bonded concrete systems: an interface fracture
approach, Ph.D. dissertation, Department of Civil and Environmental Engineering,
Massachusetts Institute of Technology, February 2005, Available through URL:
/https://dspace.mit.edu/bitstream/1721.1/28932/1/60525855.pdfS.
[6] L. Guoqiang, S. Hedlund, S.S. Pang, W. Alaywan, J. Eggers, C. Abadie, Repair of
damaged RC columns using fast curing FRP composites, Composites: Part B 3
(2003) 261–271.
[7] J.M. Gilstrap, C.W. Dolan, Out-if plane bending of FRP-reinforced masonry
walls, Composites Science and Technology 58 (1998) 1277–1284.
[8] S. Hamoush, M. McGinley, P. Mlakar, M. Terro, Out-of-plane behavior of
surface-reinforced masonry walls, Construction and Building Materials 16 (6)
(2002) 341–351.
[9] F. Bastianini, A. Di Tommaso, G. Pascale, Ultrasonic non-destructive assessment of bonding defects in composite structural strengthening, Composite
Structures 53 (4) (2001) 463–467.
[10] M.R. Vwalluzzi, D. Tinazzi, C. Modena, Shear behavior of masonry panels
strengthened by FRP laminates, Construction and Building Materials 16 (7)
(2002) 409–416.
[11] A. Cecchi, G. Milani, A. Tralli, In-plane loaded CFRP reinforced masonry walls:
mechanical characteristics by homogenization procedures, Composites
Science and Technology 64 (13–14) (2004) 2097–2112.
[12] O. Hag-Elsafi, S. Alampali, J. Kunin, Application of FRP laminates for
strengthening of a reinforced concrete T-beam bridge structure, Composite
Structures 52 (2001) 453–466.
791
[13] J. Philip, D. De Vries, Moisture movement in porous materials under
temperature gradients, Transactions American Geophysical Union 38
(1957) 222–232.
[14] A. Luikuv, Heat and Mass Transfer in Capillary Porous Bodies, Pergamon,
Oxford, 1966.
[15] W.J. Chang, C.I. Weng, An analytical solution to coupled heat and moisture
diffusion transfer in porous materials, International Journal of Heat and Mass
Transfer 43 (2000) 3621–3632.
[16] W.J. Chang, C.I. Weng, An analytical solution of a transient hygrothermal
problem in an axisymmetric double-layer annular cylinder by linear theory
of coupled heat and moisture, Applied Mathematical Modeling 21 (1997)
721–734.
[17] R.N. Pandey, S.K. Pandey, Complete and satisfactory solutions of Luikov
equations of heat and moisture transport in a spherical capillary porous body,
International Communications in Heat and Mass Transfer 27 (No. 7) (2000)
975–984.
[18] N. Mendes, P.C. Philippi, R. Lamberts, A new mathematical method to solve
highly coupled equations of heat and mass transfer in porous media,
International Journal of Heat and Mass Transfer 45 (2002) 509–518.
[19] N.E. Wijeysundera, B.F. Zheng, M. Iqbal, E.G. Hauptmann, Numerical
simulation of the transient moisture transfer through porous
insulation, International Journal of Heat and Mass Transfer 39 (No. 5) (1995)
995–1004.
[20] V.P. De Freitus, V. Abrantes, P. Crausse, Moisture migration in building wallsanalysis of the interface phenomena, Building and Environment 31 (No. 2)
(1996) 99–108.
[21] J. Banaszak, S.J. Kowalski, Drying induced stresses estimated on the base
of elastic and viscoelastic models, Chemical Engineering Journal 86 (2002)
139–143.
[22] K. Liao, Y.M. Tan, Influence of moisture-induced stress on in situ fiber
strength degradation of unidirectional polymer composite, Composites: Part
B 32 (2001) 365–370.
[23] M.M. Abel Wahab, A.D. Crocombe, A. Beevers, K. Ebtehaj, Coupled stressdiffusion analysis for durability study in adhesively bonded joints, International Journal of Adhesion & Adhesives 22 (2002) 61–73.
[24] W. Obeid, G. Mounajed, A. Alliche, Mathematical formulation of thermohygro-mechanical coupling problem in non-saturated porous media,
Computer Methods in Applied Mechanics and Engineering 190 (2001)
5105–5122.
[25] M. Khoshbakht, M.W. Lin, C.A. Feickert, A finite element model for
hygrothermal analysis of masonry walls with FRP reinforcement, Finite
Elements in Analysis and Design 45 (8 & 9) (2009) 511–518.
[26] Moisture Analysis and Condensation Control in Building Envelopes, ASTM
Manual 40, American Society for Testing and Materials, 2001.
[27] M.W. Lin, J.B. Berman, M. Khoshbakht, C.A. Feickert, A.O. Abatan, Modeling of
moisture migration in an FRP reinforced masonry structures, Building and
Environment 41 (2006) 646–656.
[28] C.A. Feickert, and J.B. Berman, Some considerations effecting the modeling
of moisture percolation through common porous building materials.
Report, Construction Engineering Research Laboratory, Champaign, IL,
2002.
[29] FEMLAB 2.3 User’s Guide and Introduction, 2003, COMSOL AB.
[30] P.N. Brown, A.C. Hindmarsh, L.R. Petzold, Using Krylov methods in the
solution of large-scale differential-algebraic systems, SIAM Journal on
Scientific Computing 15 (1994) 1467–1488.
Téléchargement