Sunteți pe pagina 1din 19

EverFE Theory Manual

Bill Davids, Ph.D., P.E.


University of Maine
Dept. of Civil and Environmental Engineering

February 2003

EverFE version 2.23
1
1. Introduction
EverFE (current version 2.23) is a 3D finite-element analysis tool for simulating the response of
jointed plain concrete pavement systems to axle loads and environmental effects. EverFE couples a highly
interactive graphical user interface for model development and result visualization written in
Tcl/Tk/Tix/vTk with finite-element code written in object-oriented C++.
Some significant features of EverFE include:
The ability to model 1, 2, or 3 slab and/or shoulder units longitudinally and/or transversely (up to
9 slab-shoulder units total in a 3x3 configuration). Transverse tie bars between adjacent slab-
shoulder units can be explicitly modeled.
Up to three elastic base layers with a bonded or unbonded base can be specified. Slab-base shear
transfer can be captured via an elastic-plastic distributed horizontal stiffness between the slabs
and base. A tensionless or tension-supporting dense liquid foundation underlies the bottom-most
model layer.
Linear or nonlinear aggregate interlock shear transfer can be simulated at transverse joints.
Dowels can be precisely located across transverse joints, and dowel looseness modeled. In lieu of
modeling dowel looseness, a dowel-slab support modulus can be specified to model dowel-slab
interaction.
Dowel misalignment and mislocation can be modeled.
A variety of different axle configurations can be easily defined with a minimum amount of input.
Linear, bilinear, and trilinear thermal gradients through the slab thickness can be captured. This
allows the simulation of thermal effects as well as slab shrinkage.
EverFEs extensive post-processing capabilities permit the visualization of stresses,
displacements, and internal dowel forces and moments. Critical response values at any point in
the model can be easily retrieved.
This manual details the finite-element implementation of EverFE, and includes descriptions of critical
features and references to appropriate literature that augment the material presented here. For specific
instructions on the use of the software, see the help file EverFE_help.chm, which is integrated with the
Help menu in the EverFE interface. This manual is organized into the following topics: Finite-Element
Discretization; Modeling of Dowels and Ties; Aggregate Interlock Modeling; Treatment of Axle and
Thermal Loads; Finite Element Solution Strategies; and Program Architecture and File Structure.
2. Basic Finite-Element Discretization
There are five elements in EverFEs finite-element library: 20-noded quadratic brick elements are
used to discretize the slab and elastic base and sub-base layers; 8-noded planar quadratic elements
incorporate the dense liquid foundation below the bottom-most elastic layer; 16-noded quadratic interface
elements implement both aggregate interlock joint shear transfer and shear transfer at the slab-base
interface; and 3-noded embedded flexural elements are coupled with conventional 2-noded shear beam
elements to model dowels at transverse joints and ties at longitudinal joints. Figure 1 shows a finite-
element mesh of a four-slab model and the corresponding elements. The flexural elements used to model
dowels and ties are detailed separately in Section 4, and aggregate interlock modeling is covered in
Section 5.


2


Figure 1: Basic Finite-Element Discretization

2.1 Model boundary conditions
The boundary conditions are the minimum required to prevent rigid-body motion, but differ slightly
depending on whether or not an elastic base layer is explicitly modeled. In the case where a base layer is
modeled, the slabs are restrained in the horizontal x-y plane by the shear stiffness of the slab-base
interface as discussed in Section 3, and receive vertical support from contact with the base. Rigid body
motion of the base and sub-base layers is prevented by restricting the x- and y- displacements of one node
on the x face, and restricting the x-displacement of a second node on the x face. Vertical support of the
entire system is provided by the dense liquid foundation, which is always incorporated below the bottom-
most layer of the model.
If the slabs are founded directly on a dense liquid, i.e. no base layer is modeled, each slab is
restrained against x- and y-direction displacements at one node on its x face, and against x-direction
displacement at a second node on its x face to prevent rigid-body motion of each slab. Again, vertical
support is provided by the dense liquid foundation.
2.2 Modeling of the slab, base and sub-base layers
In all EverFE models, the slab, base and sub-base layers are treated as 3D, linearly elastic, isotropic
continua. Each layer is discretized with standard 20-noded serendipity brick elements. The finite-
element meshes are rectilinear, and the same number of element divisions is used for each slab and the
base/sub-base layers below the slab in the x-y plane to ensure compatibility at the slab-base interface.
Details of the brick element formulation and implementation can be found in finite-element texts
such as Zienkiewicz and Taylor (1994). To maintain generality, an isoparametric element formulation is
20-noded brick element
zero
thickness
16-noded interface element
8-noded dense liquid element
x
y
z
20-noded brick element
zero
thickness
16-noded interface element
8-noded dense liquid element
x
y
z
3
used and all required element integration is performed numerically using 8-point (2x2x2) Gauss
quadrature. The initial public release of EverFE (version 1.02, released January 1998) used 27-point
(3x3x3) Gauss integration; however, subsequent internal studies showed that the higher-order integration
scheme added to the computational time without significantly improving accuracy.
2.3 Modeling of the dense liquid foundation
The dense liquid foundation can either support tension, or be tensionless. It is important to note that
the tensionless dense liquid does not account for any pre-compression due to dead load, i.e. the total
vertical deflection including the effect of dead load must be overcome before the dense liquid foundation
stress and stiffness become zero.
The 8-noded element illustrated in Figure 1 is used to discretize the dense liquid. This element was
formulated specifically for this application, and full details of the implementation can be found in Davids
(1998). The element incorporates standard quadratic shape functions for interpolation of vertical
displacements within the element (Zienkiewicz and Taylor 1994), ensuring that it displaces compatibly
with the 20-noded brick element with which it shares nodes. An isoparametric element formulation is
used, and all necessary element integrations are performed numerically using 9-point (3x3) Guass
quadrature to ensure accurate results when the tensionless option is selected.
The only constitutive parameter needed for this element is the distributed stiffness of the dense liquid
foundation [force/volume]. For the tensionless foundation, if tension occurs at an element integration
point during the solution process, the stress and stiffness at that point are set to zero during integration of
the element stiffness matrix and equivalent force vector. For the conventional, tension-supporting dense
liquid, the stiffness remains constant at all points.
3. Treatment of the Slab-Base Interface
Modeling interaction of the slab and base is crucial to predicting pavement response to axle loads
near joints and to thermal or shrinkage gradients. EverFE allows the consideration of either perfect bond
between the slab and base, or separation of the slab and base under tension. In both cases, the slab and
base do not share nodes, and nodal constraints are used to satisfy the required contact conditions (see
Figure 2). The solution algorithm relies on a perturbed Lagrangian formulation and a nodal constraint-
updating scheme based on the current normal stress between the slab and base. More details on the global
nonlinear solution strategy are given in Section 7.
Shear transfer between the slab and base can be important when analyzing pavements subjected to
thermal and/or shrinkage strains. Rasmussen and Rozycki (2001) overviewed the factors governing slab-
base shear transfer, noting that both friction and interlock between the slab and base play a role. In
addition, they calibrated a bilinear, elastic-plastic shear transfer model from results of push tests of slabs
Base element Interface
elements transfer
shear stress
Slab element
Pairs of nodes vertically constrained
if compression at interface
x or y
z
x or y
k
SB
o
o
Interface constitutive relationship
Figure 2: Modeling Separation and Shear Transfer at the Slab-Base Interface
Base element Interface
elements transfer
shear stress
Slab element
Pairs of nodes vertically constrained
if compression at interface
x or y
z
x or y
k
SB
o
o
Interface constitutive relationship
Figure 2: Modeling Separation and Shear Transfer at the Slab-Base Interface
4
on various base types. One conclusion of the their study was that the effect of slab-base shear transfer
should be incorporated in 3D analyses of pavement systems. Another study by Zhang and Li (2001)
focused on developing a one-dimensional analytical model for predicting shrinkage-induced stresses in
concrete pavements that accounts for slab-base shear transfer. Like the model developed by Rasmussen
and Rozycki, their model ultimately relied on a bilinear, elastic-plastic shear transfer model. Zhang and Li
concluded that the type of supporting base and thus the degree to which it restrains slab shrinkage
significantly affects slab stresses.
To capture slab-base shear transfer, EverFE employs 16-noded, zero-thickness quadratic interface
elements that are meshed between the slab and base as shown in Figures 1 and 2. The element
incorporates standard quadratic shape functions for interpolation of displacements (Zienkiewicz and
Taylor 1994), ensuring that it displaces compatibly with the 20-node brick elements with which it shares
nodes. The element tracks relative displacements between the slab and base in the vertical (z) and both
horizontal (x and y) directions. An isoparametric element formulation is used, and all necessary element
integrations are performed numerically using 9-point (3x3) Guass quadrature.
The bilinear element constitutive relationship is based on that given by Rasmussen and Rozycki
(2001) and Zhang and Li (2001). Figure 2 illustrates the constitutive relationship, which is characterized
by an initial distributed stiffness k
SB
|force/voIume| and sIip dispIacement
o
. While k
SB
has the same units
as the well-known dense liquid foundation modulus, k
SB
is a distributed stiffness in the x- and y-directions,
and the shear stresses in the x-y plane at the slab-base interface are caused by relative horizontal
displacements between the slab and base layer. This constitutive relationship is assumed to apply
independently in both the x- and y-directions as long as the slab and base remain in contact (i.e. a
compressive normal stress exists at the slab-base interface). The fact that there will be little or no shear
transfer when slab-base separation occurs is accommodated by setting the interface stiffness and shear
stress to zero whenever the reIative verticaI dispIacement z > 0. Modeling this loss of shear transfer with
loss of slab-base contact can be important, especially when thermal gradients are simulated (Davids et al.
2003).
Note that unlike a frictional model, the shear stress does not depend on the magnitude of the normal
stress. However, for very large values of k
SB
, this model approaches Coulomb friction with a very large
friction coefficient, and for very small values of k
SB
, it is equivalent to a frictionless interface. An
advantage of the modeling scheme used by EverFE is that the symmetry of the system stiffness equations
is maintained, which allows the use of the highly efficient preconditioned conjugate-gradient solver
discussed in Section 7. Idealizing slab-base interaction with conventional Coulomb friction would destroy
this symmetry, requiring the use of more complex (and likely less efficient) solution techniques.
4. Modeling of Dowels and Ties
EverFE models dowels and transverse tie bars explicitly with 3-
noded, quadratic embedded flexural finite elements (Davids and
Turkiyyah 1997; Davids 2000). This approach has the advantage of
allowing the dowels and tie bars to be precisely located irrespective
of the slab mesh lines as shown in Figure 3. This embedded element
formulation also permits significant savings in computation time by
allowing a range of load transfer efficiencies to be simulated without
requiring a highly refined mesh at the joint. Additionally, the
embedded dowel formulation allows the rigorous treatment of gaps
between the dowels and the slabs (dowel looseness). Alternatively
the dowels can be modeled as beams on an elastic foundation, where
Winkler foundation springs are sandwiched between the dowels and
the slabs. These two options for capturing dowel-slab interaction are
shown conceptually in Figure 4(a).
embedded dowel
element
solid
element
Figure 3: Dowel Element
embedded dowel
element
solid
element
Figure 3: Dowel Element
5

4.1 Embedded Dowel Element Formulation
The finite-element formulation of the embedded element relies on expressing the nodal displacements of
the dowel as functions of the nodal displacements of the solid element it is embedded within. The three-noded
quadratic beam element (that originally has 18 degrees-of-freedom in its element displacement vector, U
d
)
takes on the degrees-of-freedom of the solid element it is embedded within, U
e
. The element displacement
vector of the embedded bending element, U
de
, becomes:
) 1 (

=
e
d
de
U
U
U
As shown in Davids and Turkiyyah (1997), U
d
can be recovered through the following matrix
transformation:
) 2 (
de d
TU U =
The matrix T contains shape functions of the embedding element and constrains the dowel to the
embedding element. It follows that the tangent stiffness matrix of the embedded element K
de
, needed for
the nonlinear solver, can be determined from the original dowel element stiffness, K, as follows:
) 3 ( KT T K
T
=
de

If dowel looseness is explicitly modeled, a nodal contact approach is employed where T encapsulates
all the necessary information regarding constraints. The advantage of this approach is that the contact
nonlinearity resulting from the rigorous simulation of gaps between the dowels and slabs is treated like a
x
x
y
Plan View

Original
position
Misaligned
position
x
z
Elevation View
z
Original
position
Misaligned
position

(b) Dowel Misalignment


q
r
q s
Slab
Dowel-slab springs
C.L. joint
(a) Dowel-Slab Interaction
Gap between
dowel and slab
Figure 4: Dowel Modeling
x
x
y
Plan View

Original
position
Misaligned
position
x
z
Elevation View
z
Original
position
Misaligned
position

(b) Dowel Misalignment


q
r
q s
Slab
Dowel-slab springs
C.L. joint
(a) Dowel-Slab Interaction
Gap between
dowel and slab
x
x
y
Plan View

Original
position
Misaligned
position
x
z
Elevation View
z
Original
position
Misaligned
position

(b) Dowel Misalignment


q
r
q s
Slab
Dowel-slab springs
C.L. joint
(a) Dowel-Slab Interaction
Gap between
dowel and slab
Figure 4: Dowel Modeling
6
material nonlinearity through the simple stiffness matrix transformation of Equation 3. Internally, the
contact conditions are checked and updated at each iteration of the nonlinear solver.
With the foregoing formulation, the embedded bending element has also been extended to permit the
inclusion of general bond-slip relations between the dowel and surrounding slab (Davids 2000). If the
incremental vector of relative displacements between the slab and dowel is denoted as d , and it is
assumed that the corresponding incremental force vector df can be computed as:
) 4 ( = d d D f
In Equation 4, D is a 3x3 constitutive matrix. It has been shown that the stiffness matrix of the embedded
dowel element with general bond-slip relations, K
dt
, can then be computed as:
) 5 (
1
1

+ = hd
de dt
DB B K K
T

In Equation 5, B is a matrix operator containing shape functions of both the embedding solid element and
the dowel element, and h is the length of the dowel element; integration is performed with respect to the
dowel element local coordinate, .
Physically, this formulation of the embedded element with a general bond-slip relationship between
the dowel and slab is analogous to the classic beam on elastic foundation, but differs in that forces in the
three coordinate directions can exist between the dowel and the slab. The magnitude of the forces depends
on the relative displacement between the dowel and the slab and the constitutive relations of the
dowel/slab interface incorporated in the matrix, D.
4.2 Implementation in EverFE
In EverFE, the 3-noded quadratic embedded dowel elements are used to model the portions of the
dowels embedded in the slabs. To ensure accurate results, 12 embedded elements are used to model the
embedded portion of the dowels on each side of each transverse joint, and a 2-noded shear beam is used
to model the portion of the dowel spanning the joint.
When dowel looseness is modeled, 10 of the 12 embedded dowel elements are used over the portion
of the dowel where there is a gap between the dowel and slab to ensure a sufficient number of potential
points of contact between the dowels and slabs. If the dowel is treated as a beam on a dense liquid
foundation, the 12 elements are evenly distributed along the embedded portion of the dowel. The diagonal
terms of D corresponding to the model y- and z- coordinates, D(2,2) and D(3,3), are the dowel-slab
support modulus specified in EverFE. The dowel-slab support modulus is computed as the product Kd,
where K is the commonly used modulus of dowel support [force/volume], and d is the dowel diameter.
The paper by Dei Poli et al. (1992) discusses the basis for the development of K from the properties of the
slab concrete and dowel; additionally, the EverFE help manual includes the results of a parametric study
showing the effect of the parameter Kd on load transfer efficiency for a simple two-slab model. D(1,1)
applies in the x-direction, and is the dowel-slab restraint modulus that controls the degree of bond
between the dowels and slabs. A large value of the dowel-slab restraint modulus implies a high degree of
bond between the dowels and slabs, and a value of 0 implies no bond. For an example illustrating the
effect of this parameter, see Davids et al. (2003).
Transverse ties are modeled in the same manner as the dowels, although only the dense liquid support
option is available, and fewer elements are used to discretize the ties since their shears and moments are
of secondary interest.
4.3 Simulation of Dowel Misalignment/Mislocation
EverFE also allows the simulation of dowel misalignment and/or mislocation through the specification of
four parameters ( x, z, , ) that shift an individuaI doweI aIong the x- and z-axes and define its angular
misalignment in the horizontal and vertical planes (see Figure 4(b)). The dowel support and restraint moduli
coincide with the local dowel coordinate axes (q,r,s), which are rotated from the global (x,y,z) axes by the
7
angIes and . The meshing aIgorithm preciseIy Iocates individuaI fIexuraI eIements within the soIid
elements by first solving for the intersection of each dowel with solid element faces, and then subdividing
each dowel into at least 12 individual quadratic embedded flexural elements on each side of the joint face as
discussed previously.
5. Aggregate Interlock Modeling
Aggregate interlock shear transfer is assumed to occur across the entire width of each transverse joint
in the finite-element model. Both linear and nonlinear options are available for modeling aggregate
interlock. With the linear option, the shear stress developed between the joint faces is proportional to the
relative vertical movement at the joint, and the shear stress is independent of the joint opening. The
nonlinear option includes both the nonlinearity in the shear stress-relative vertical displacement relation
as well as the nonlinear variation in shear stress transfer with changes in joint opening. The basis for and
implementation of both of these options is detailed in this section.
In both cases, EverFE employs a 16-noded, zero-thickness quadratic interface element that is meshed
between the joint faces as shown in Figure 1. This is the same element detailed in Section 3 of this
manual that is used to capture shear transfer at the slab-base interfaces. As discussed in Section 3, an
isoparametric element formulation is used, and all necessary element integrations are performed
numerically using 9-point (3x3) Guass quadrature.
5.1 Linear Aggregate Interlock Load Transfer
The linear option is the simplest approach for modeling aggregate interlock load transfer at
longitudinal joints. In this case, only a joint stiffness is specified to control the degree of aggregate
interlock load transfer. The units of this parameter are force/volume, and it is analogous to a dense liquid
k in that it can be interpreted as a spring stiffness per unit area.
The specified joint stiffness is constant over the entire area of the joint, and does not vary with
relative vertical displacement or joint opening. If the joint stiffness is set to zero (the default value), there
will be no aggregate interlock load transfer, and a very large value will result in high load transfer
efficiency. The joint stiffness applies only in the vertical (z) direction, and y-direction relative joint
movement is unrestrained.
It is worthwhile noting that a number of prior studies have used linear springs to model aggregate
interlock load transfer across pavement joints (Ioannides and Korovesis 1990; Kuo et al. 1995; Brill et al.
1997). Further, dowel load transfer has also been idealized with linear springs spanning transverse joints
(Ionnides and Korovesis 1992). To examine the accuracy and usefulness of this approach, consider the
following example consisting of a simple two-slab (one row, two column) model with a 250 mm thick
slab (E = 28000 MPa, = 0.20, density = 0) founded directly on a dense liquid with k = 0.03 MPa/mm.
Two cases have been analyzed: the first considers only linear aggregate interlock load transfer, with joint
stiffnesses ranging from 0 to 10 MPa/mm; the second considers no aggregate interlock load transfer, but
has 11-32 mm dowels spaced at 300 mm on center with dowel-slab support moduli ranging from 0 to
100,000 MPa. In both cases, the slab is subjected to an 80 kN axle located at the joint face. Each slab is
meshed with 12x12x2 elements. Figure 5 shows a picture of the doweled model.
After running multiple models with varying aggregate interlock joint stiffnesses and dowel-slab
support moduli, joint load transfer efficiencies (LTEs) were computed and peak slab tensile stresses under
each wheel were tabulated. Figure 6 shows the variation in peak tensile stress with LTE for both joint
stiffness and dowel-slab support modulus. Note how the doweled model consistently predicts a lower slab
stress for a given load transfer efficiency, except at the extreme values of 100% and 0% LTE. This can be
attributed to the fact a dowel falls directly below each wheel (see Figure 5), providing a concentrated
source of joint load transfer, whereas the constant aggregate interlock joint stiffness is evenly distributed
across the joint width. When there is perfect or no load transfer, the localized nature of the dowel support
has no effect on slab stresses.
8

Figure 5: EverFE Screen Shot of Aggregate Interlock Example
Figure 6: Variations in Slab Stress with Load Transfer Efficiency

0 10 20 30
40 50 60 70 80 90 100
0.9
1
1.1
1.2
1.3
1.4
1.5
Evenly Spaced Dowels
Linear Aggregate Interlock
Load Transfer Efficiency (%)
P
e
a
k

S
l
a
b

S
t
r
e
s
s

(
M
P
a
)

9
5.2 Nonlinear Aggregate Interlock Load Transfer
The nonlinear aggregate interlock load transfer option allows the consideration of both the effect of
relative vertical joint displacement and joint opening on aggregate interlock load transfer effectiveness.
EverFE relies on a two-phase aggregate interlock model developed by Walraven (1981, 1994) to generate
nonlinear aggregate interlock crack constitutive relations. The crack is assumed to follow the aggregate
particle boundaries, and the aggregate particles bearing on the cement paste are taken to be at the point of
slip. Walravens model also assumes that the aggregate particles are graded according to a Fuller
distribution, and the maximum particle diameter, D
max
, and the aggregate volume fraction, p
k
, are model
parameters.
Given an ultimate strength of the cement paste,
pu
, a coefficient of friction between the paste and
aggregate of , and computing the projected contact areas between the aggregate and paste using the
deformed geometry, the forces required for equilibrium of a single aggregate particle
diameter/embedment combination can be computed. Using the probability of occurrence for a particular
embedment/diameter combination derived by Walraven (1981), the likely forces on all particles are then
summed to give the total forces acting on a crack plane for a given relative slip displacement and joint
opening.
Typical crack shear stress-displacement relations predicted by the model are shown in Figure 7,
where each curve corresponds to a specific joint opening. Although only 3 curves are shown, EverFE
internally generates 40 curves over a range of joint openings between 0 and 20 mm to give a very
complete definition of the shear stress-displacement relations. The majority of these curves apply for joint
openings between 0 and 2 mm, where the most rapid changes in load transfer effectiveness with joint
opening occur. As with the linear model, shear stress is transferred only in the vertical direction.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
1
2
3
4
5
6
7
8
9
Relative Vertical Joint Displacement (mm)
S
h
e
a
r

S
t
r
e
s
s

(
M
P
a
)

0.02 mm Joint Opening
0.12 mm Joint Opening
0.98 mm Joint Opening

Figure 7: Typical Nonlinear Aggregate Interlock Shear Constitutive Relations

10
The implementation of Walravens model in EverFE is consistent with the general formulation of a
materially nonlinear finite-element analysis. The necessary tangent moduli are computed numerically
from the constitutive relations, and stresses are numerically interpolated from the constitutive relations for
a given joint opening and relative vertical displacement. It is important to note that the finite-element
implementation accounts for the variation in joint opening with vertical position on the joint face that
develops under loading of the pavement system. One limitation of the model, however, is that it does not
account for the sawcut at the top of a typical contraction joint where no aggregate interlock load transfer
would normally occur.
The parameters necessary to define the nonlinear aggregate interlock model in EverFE are the
maximum paste strength,
pu
, the paste-aggregate coefficient of friction , the aggregate voIume fraction,
p
k
, and the maximum aggregate diameter, D
max
. Walraven (1994) suggested the following relationship
between the compressive strength of a 150 mm concrete cube, f
cc
, and
pu
, where both are in MPa:
) 6 ( 0 . 8
cc pu
f =
The compressive strength of a 150 mm concrete cube can be assumed to be approximately 1.25 times the
strength of a standard concrete cylinder,

c
f (Wang and Salmon 1985). In addition, when weak aggregate
is used that is prone to fracture, Walraven suggested proportioning
pu
downward by a fracture index, C
f
<
1.0. Recent research on the topic of aggregate interlock joint shear transfer (Jensen and Hansen 2003;
Wattar et al. 2001) suggests that the basic concepts underlying Walravens model are sound. However, its
accuracy in predicting pavement joint load transfer may vary with specific characteristics of aggregate
shape and type, as well as the degree of damage at the joint. In addition, a study by Davids and Mahoney
(1999) has shown good qualitative and reasonable quantitative agreement between existing experimental
data and predictions of aggregate interlock load transfer efficiency using Walravens model.
6. Treatment of Axle Loads and Thermal Effects
EverFE allows the consideration of simultaneous axle loads and prestrains due to thermal or
shrinkage effects. This section documents the methods by which these loads are included in each EverFE
finite-element model.
6.1 Axle Loads
The complex axle configurations available in EverFE are simply collections of single rectangular
wheel loads, and each wheel is treated identically by the finite-element code. A wheel load is defined by
the (x,y) location of its geometric center, the length L and width W of the tire contact area, and the
magnitude of the wheel load P. The load is assumed to produce a constant pressure over the wheel contact
area.
The critical issue regarding the application of the wheel loads in the finite-element model is
determining the set of nodal forces that are equivalent to the uniformly distributed pressure generated by
the wheel. This is challenging since the wheel load contact area is not restricted to coincide with an
element face, and in fact can partially load several element faces. EverFE handles this by dividing each
wheel contact area into smaller rectangular sub-areas by using a grid having n
x
x n
y
divisions along each
edge. The i
th
sub-area of the wheel defined by the grid thus has an area of LW/(n
x
n
y
) and sees a total force
of p
i
= P/(n
x
n
y
).
The equivalent nodal force vector due to each p
i
is then computed by first determining the solid
element that it contacts using the same fast geometric search procedures needed for the finite-element
solver that is discussed in Section 7. The work-equivalent set of nodal forces is then computed as the
product of p
i
and the vector of element shape functions evaluated at the point of application of p
i
. The
sum of all work-equivalent nodal force vectors is the total nodal force vector applied by the entire wheel.
This procedure is consistent with the virtual work and energy principles that form the basis of the finite-
11
element method, using a rectangular rule to numerically integrate the tire contact pressure over its area of
application.
Clearly, the critical parameters in this calculation are n
x
and n
y
. To examine their effect on solution
accuracy, consider the following example where a single slab founded on a dense liquid is subjected to a
single 40 kN wheel load at its edge with L = W = 200 mm. The slab is 4400 mm long, 3600 mm wide, and
254 mm thick with E = 27,600 MPa and 0.20. The sIab is meshed with 12 x 12 eIements in pIan and 2
elements through the slab thickness, and is shown in Figure 8. The model was solved using values of n
x
=
n
y
ranging between 1 and 20. When n
x
= n
y
= 1, the wheel load is treated as a single point load. Table 1
gives the x-direction stresses at the top and bottom of the slab for different values of n
x
= n
y
, and shows
clear convergence by n
x
= n
y
= 10. Since the equivalent nodal force calculation presented here is not
computationally expensive, EverFE uses a fixed value of n
x
= n
y
= 20 for all simulations to ensure
accuracy for a wide range of wheel load sizes and levels of mesh refinement. It should be noted that other
researchers have developed different methods of handling the problem of wheel load contact stresses
(Hjelmstat et al. 1997); however, the method presented here is conceptually simple, easily implemented,
and accurate.
Difficulties can arise when one or more of the sub-area loads falls outside the slab boundary, and
cannot be located in an element. EverFE handles this by moving the point of application of p
i
around a
circle with an ever-increasing radius until p
i
falls within a solid element, and the calculation then proceeds
as detailed. This may degrade the accuracy of the method, however, and wheel loads having portions of
their contact areas outside slab boundaries should be avoided.




Figure 8: Plan View of Mesh for Wheel Load Refinement Study

12
Table 1: Effect of n
x
and n
y
on Slab Stresses

n
x
, n
y

Top of Slab
Stress (MPa)
Bottom of Slab
Stress (MPa)
1, 1 -2.13 1.79
2, 2 -1.78 1.67
3, 3 -1.81 1.68
4, 4 -1.78 1.67
5, 5 -1.79 1.67
10, 10 -1.77 1.67
15, 15 -1.77 1.67
20, 20 -1.77 1.67

6.2 Thermal and Shrinkage Effects
Thermal and shrinkage effects are treated as general element prestrains in a manner consistent with
fundamental finite-element theory (Zienkiewicz and Taylor 1994). EverFE allows the specification of
linear, bi-linear, or tri-linear temperature changes throughout the slab thickness, and the prestrain is
computed as the product of the specified temperature change and the user-specified coefficient of thermal
expansion. Specification of shrinkage strains can be accomplished by converting the desired shrinkage
strain into an equivalent temperature change using the coefficient of thermal expansion. The element
prestrains are converted to an equivalent nodal force vector by the usual element integration, and are
subtracted from the total strain during the calculation of internal stresses.
An important detail that must be emphasized is that the 20-noded brick elements used by EverFE are
capable of capturing an essentially linear variation in strain over their volume. Hence, when a bi-linear
thermal gradient is specified, the finite-element mesh must have an even number of elements through its
thickness to ensure accurate prediction of stresses and displacements. Similarly, if a tri-linear thermal
gradient is specified, there must be 3,6,9, etc. elements through the slab thickness. These mesh restrictions
are automatically enforced by the EverFE.
7. Finite Element Solution Strategies
EverFE was originally developed for use on desktop computers, and as a result uses solution
strategies designed to minimize computational time and memory usage in its complex 3D finite element
analyses. At the core of the solver for both linear and nonlinear problems is the solution of a system of
symmetric, positive-definite linear equations. To accomplish this, EverFE relies on an iterative multigrid-
preconditioned conjugate gradient (MG-PCG) solver developed specifically for use on 3D finite element
models that incorporate the multiple element types and the phenomena detailed in this manual. This
section overviews EverFEs global nonlinear solution strategy, and provides information on the core MG-
PCG solver. For more details, see Davids and Turkiyyah (1999).
7.1 Global Nonlinear Solution Strategy
EverFE finite-element models can be either linear or nonlinear. Nonlinearity is due to either material
properties (use of a tensionless dense liquid foundation, dowel looseness, the nonlinear aggregate
interlock model, or non-zero slab-base shear transfer when the slab-base interface is unbonded) or contact
conditions arising from an unbonded slab-base interface. As noted in Section 4, dowel looseness is
actually a contact nonlinearity that is more conveniently treated as a material nonlinearity.
The nonlinear solution strategy used by EverFE is essentially a full Newton method with the updating
of contact constraints performed at each Newton iteration. The procedure is given in Algorithm 1, where
K
k
is the tangent system stiffness matrix at the k
th
iteration; dU
k
is the update to the system displacement
vector, U; F is the vector of applied forces; and r is the residual vector of unbalanced forces. A solution is
13
achieved when r is sufficiently close to zero. If the model is linear, K
k
is constant, and only a single
iteration is required.
r = F
while ||r||/||F|| > 10
-04

update contact constraints
update K
k

solve K
k
dU
k
= r
U
k
= U
k-1
+ dU
k

update r
end
Algorithm 1: Nonlinear Solution Strategy

It is worth noting that for models with a base layer, the solution of K
k
dU
k
= r for dU
k
involves a sub-
iteration using a technique called Uzawas method to satisfy the contact constraints that is not detailed
here. This is necessary to avoid ill conditioning of K
k
and maintain the efficiency of the MG-PCG solver.
For nonlinear problems, the contribution of the 20-noded solid elements used to discretize the linearly
elastic slabs and base to K
k
is not updated. This saves significant computational time, since much of the
work involved in forming K
k
arises from the numerical integration of these large element stiffness
matrices.

7.2 Overview of Multigrid Solver
When solving a large, 3D finite element problem either linear or nonlinear solution of the system
stiffness equation
) 7 ( r KU =
requires the bulk of the computational resources. For a nonlinear problem, K and U in Equation 7
correspond to K
k
and dU
k
in Algorithm 1. In most codes, direct solution methods such as LU factorization
are employed to solve Equation 7, where K is factored into upper and lower triangular matrices, and the
solution vector U is computed through forward and back substitution. While factorization is
straightforward and relatively insensitive to poor conditioning of K, the amount of work required to factor
K grows at least quadratically with the number of unknowns for realistic problems (even when sparse
matrices are utilized). An additional barrier to the use of LU factorization is the additional memory
required to store the matrix factors, which is significantly more than that required to store K itself for
sparse matrices.
To overcome this problem, EverFE employs a highly efficient, multigrid-preconditioned conjugate
gradient solver. The conjugate gradient method is a widely-used iterative technique for solving systems
such as Equation 7 that are characterized by a symmetric, positive-definite coefficient matrix, and the
basic algorithm is easily implemented. However, it is well known that the efficiency of the conjugate
gradient method is highly dependent on the efficiency of the preconditioner (Saad 1996).
EverFE relies on the multigrid method to precondition the conjugate gradient method. Multigrid
methods are themselves iterative techniques for the solution of discretizations of partial differential
equations that rely on multiple discretizations of the same domain (Brandt 1977). Denoting the current
error in the solution vector by e, the vector of residual nodal forces as r = F KU, and the exact
(unknown) solution as U
*
, we can write:
14
) 9 (
) 8 (
*
r Ke
U U e
=
=

A small number of Gauss-Seidel iterations are performed for the finer meshes to remove high-
frequency error components, and the low frequency error components are approximated via a direct and
inexpensive solution on the coarsest mesh (Brandt 1977). Figure 9 presents the algorithm and a
conceptual overview for a two-mesh sequence. EverFE relies on a V-cycle multigrid scheme where r is
sequentially restricted from the finest to the coarsest mesh with symmetric Gauss-Seidel smoothing
performed at each step. Following the solution on the coarsest grid, the approximated error vector is
sequentially interpolated and smoothed from the coarsest to the finest mesh. EverFEs use of a single
multigrid V-cycle to precondition a conjugate gradient iteration takes advantage of the symmetric positive
definiteness of the system stiffness equations.
One of the primary difficulties in implementing a multigrid scheme for un-nested mesh sequences of
spatially inhomogeneous domains such as those found in layered foundations is defining appropriate
restriction and interpolation operators. Restriction can be viewed as computing a force vector on a coarse
mesh that is statically equivalent to the known force vector on a finer mesh. This process is often
expressed in matrix form as follows, where r
c
denotes a coarse mesh residual force vector, r
f
denotes the
known fine mesh force vector, and R is the restriction operator:
) 10 (
f c
Rr r =
Similarly, interpolation is defined as the process of approximating the fine mesh error in the displacement
vector, e
f
, from a known coarse mesh error, e
c
using the interpolation operator, T:
) 11 (
c f
Te e =

The multigrid implementation used by EverFE relies on the element shape functions to define R and
T, which has been shown to be advantageous (Davids and Turkiyyah 1999; Fish et al. 1996). This allows
the restriction and interpolation operations to be performed on a node-by-node basis, provided that for
each fine mesh node, the coarse mesh element it lies within and the corresponding coarse mesh element
coordinates are known. Efficiently establishing this information is critical, and is achieved using
geometric search procedures detailed by Davids and Turkiyyah (1999). This general approach also
permits the easy meshing of solid and bending elements within the same model, since all calculations are
c f
Te e =
f c
Rr r =
( )
( )
f
c f
c
f c
f f f f
f
f f
smooth
smooth
subroutine
e
Te e
r K e
Rr r
e K r r
e
e r
=
=
=
=
1
) , ( MultiGrid
restrict
smooth
interpolate
Figure 9: Multigrid Concept
c f
Te e =
f c
Rr r =
( )
( )
f
c f
c
f c
f f f f
f
f f
smooth
smooth
subroutine
e
Te e
r K e
Rr r
e K r r
e
e r
=
=
=
=
1
) , ( MultiGrid
restrict
smooth
interpolate
Figure 9: Multigrid Concept
15
performed at the element level. This is critical, since the dowels are explicitly modeled as flexural
elements as discussed in Section 4.

7.2 Demonstration of Solver Efficiency
To illustrate the efficiency of EverFEs solver, consider the solution of a model of a single 4600 mm
long by 3600 mm wide by 250 mm thick slab resting directly on a dense liquid and subjected to a single
axle load. Since the model is linear, the solution of the system stiffness equations is performed only once.
Models with increasing numbers of elements and a constant maximum element aspect ratio of 4.6 were
generated and solved using both a sparse direct solver employing LU factorization and using EverFEs
MG-PCG solver. The solutions were generated on a Dell Optiplex with a 2.8MHz Pentium IV processor
and an 800 MHz front side bus. All solutions were achieved without exceeding the machines core RAM
of 1 GB. It must be noted that accounting for the symmetry of the system stiffness matrix could have
increased the efficiency of the LU factorization; however, as shown in Table 2, the results are sill
dramatic.
Table 2: Solver Efficiency Comparison

MG-PCG LU Factorization
Mesh
(n
x
x n
y
x n
z
)
Degrees-
of-
freedom
Time
(seconds)
Memory
(MB)
Time
(seconds)
Memory
(MB)
4 x 4 x 1 465 1 10 1 -
8 x 8 x 2 2,511 2 15 1 -
12 x 12 x 3 7,293 3 28 6 75
16 x 16 x 4 15,963 8 53 28 230
20 x 20 x 5 29,673 15 91 123 607
24 x 24 x 6 49,575 23 150 * *
28 x 28 x 7 76,821 39 228 * *
*Solution could not be achieved without exceeding core RAM of 1 GB
The results clearly illustrate the relative efficiency of EverFEs MG-PCG solver, especially for
medium and large-sized problems. The times reported are totals, including time required to read all input
data and write output stresses and displacements. Also of significance is nearly linear increase in solution
time and memory usage with the increase in number of unknowns (degrees-of-freedom) when the MG-
PCG solver is used. It must be noted that this linear increase in computational time cannot be expected for
all models, especially nonlinear models with either dowel looseness or slab-base contact conditions.
Finally, while the larger meshes used here are not typical for an EverFE model in that they have a large
number of elements through the slab thickness, their overall size is not unusually large. For example, the
sample project nine_slab_base that is installed with EverFE has 55,398 degrees-of-freedom, which is
larger than the single-slab mesh with 24 x 24 x 6 elements.
One disadvantage of the MG-PCG solver used by EverFE and a disadvantage with all iterative
solvers is its sensitivity to ill conditioning of K, which can arise from both the use of high values for
material stiffnesses and large element aspect ratios. For example, if a large value is specified for the
stiffness of the slab-base interface (see Section 3) or the dowel support moduli (see Section 4) the
efficiency of the solver may suffer. Keeping maximum element aspect ratios less than the suggested limit
of 5.0 easily controls the sensitivity to large element aspect ratios. If elements with aspect ratios greater
than 5.0 are used in less critical regions of the model, such as shoulders, solution time may increase.
8. Program Architecture and File Structure
EverFE consists of four separate programs: the finite-element solver, which is written in ANSI
standard, object-oriented C++; the meshing software, which is also written in C++; the C++ program that
16
generates nonlinear aggregate interlock constitutive relationships; and the user interface, which is written
in the scripting language Tcl/Tk.
Because of computational requirements, it was necessary to develop the finite-element, meshing, and
aggregate interlock model definition code using a low-level compiled language such as Fortran or C/C++.
Ultimately, C++ was chosen for two reasons: its flexibility (and thus the ease with which it allows the
implementation of the complex solution algorithms and specialized elements detailed in this manual), and
its effective dynamic memory allocation capabilities, which are crucial when solving large problems on
desktop computers. The C++ code is extensive, and details of its architecture are not presented here; for
more information, see Davids (1998).
The user interface is written in Tcl/Tk version 8.3, a freely available scripting language that runs on
multiple Unix, Windows and Macintosh platforms. Many of the higher-level interface widgets (dialog
boxes, checkboxes, menus, etc.) used by EverFE were taken from the additional Tcl/Tk package Tix
version 8.2, which is also freely available. Visualization of stresses, displacements, and dowel shears and
moments is accomplished with the Visualization Toolkit (vtk), freely available visualization software that
can be used with both Tcl/Tk and C++ (in EverFE it is called directly from Tcl/Tk). All of the EverFE
source code written in Tcl/Tk is installed with EverFE and used in uncompiled form, and these files must
never be modified.
The remainder of this section provides details on directory-file structure and the interaction of
EverFEs constituent programs.
8.1 Directory-File Structure and Program Interaction
EverFE, like most software, is installed in a user-specified directory (the default location is
C:\Program Files\EverFE2.23). Figure 10 shows EverFEs top-level directory structure.

All of the Tcl/Tk scripts are located in the scripts subdirectory. The FE-solver subdirectory
contains the files driver.exe (the executable that generates the finite-element input files) new_fe.exe
(the main finite-element executable), and agg_int.exe (a separate executable program that generates
and saves the nonlinear aggregate interlock constitutive model information used by the finite-element
code).
Basic model data is stored in the data subdirectory, which for each project contains a single file with
a .prj extension and a subdirectory that is necessary to store the project definition an analysis results.
The name of both the .prj file and the subdirectory corresponding to an EverFE project are identical to
the user-specified name the project is saved under. Within each project subdirectory, EverFE saves a file
called model_params.dat, an ASCII text file containing the information necessary to define a project
(model geometry, material properties, dowel information, meshing parameters, loading information, etc.).
Figure 10: EverFE Directory Structure
User-specified installation directory
Stores nonlinear aggregate interlock models
Stores all project definitions and results
Contains interactive help file and theory manual
Contains finite-element executables
Stores Tcl/Tk scripts
Pre-compiled Tcl/Tk/Tix/vtk libraries
Figure 10: EverFE Directory Structure
User-specified installation directory
Stores nonlinear aggregate interlock models
Stores all project definitions and results
Contains interactive help file and theory manual
Contains finite-element executables
Stores Tcl/Tk scripts
Pre-compiled Tcl/Tk/Tix/vtk libraries
17
The Tcl/Tk scripts read this file when a project is opened, and it also serves as the sole input source for
the program driver.exe, which generates the input files defining the finite-element mesh needed by
new_fe.exe. Note that the multigrid-preconditioned conjugate gradient solver detailed in Section 7
utilizes three finite-element meshes with decreasing levels of refinement, and thus driver.exe actually
generates three separate input files each time an analysis is executed. To save disk space, these files are
deleted after the analysis is completed. When an EverFE analysis executes successfully, additional files
are written to the project subdirectory that contain the model stresses, displacements and dowel results.
These output files are read directly by Tcl/Tk source code when the user enters any component of the
visualization panel; they are stored until the model is saved without re-analysis, at which point they are
deleted to ensure that a project never has a finite-element model definition that is inconsistent with the
saved output.
The agg_int directory contains input and output data for each nonlinear aggregate interlock model
that is generated and saved. There are two subdirectories within agg_int: one for models with metric
units and one for models with English units. The documentation directory contains two files:
EverFE_Help.chm, the compiled interactive help file developed with Microsofts HtmlHelp utility that
is accessed directly from the EverFE2.23 Help menu, and the file theory_manual.pdf. Finally, the
tcl_bins directory contains all of the binary code and libraries necessary to run Tcl/Tk/Tix/vtk.
When pre-defined projects are run in batch mode, the Tcl/Tk code simply puts the programs
driver.exe and new_fe.exe in a loop, sequentially analyzing each project.
9. References
Brandt, A. (1977). Multi-Level Adaptive Solutions to Boundary-Value Problems. Mathematics of
Computation, 31(138):333 390.
Brill, D.R., Hayhoe, G.F. and Lee, X. (1997). Three-Dimensional Finite-Element Modeling of Rigid
Pavement Structures. Aircraft Pavement Technology: In the Midst of Change, ASCE, pp. 151-165.
Davids, W. and Turkiyyah, G. (1997). Development of Embedded Bending Member to Model Dowel Action.
Journal of Structural Engineering, ASCE, 123(10):1312 1320.
Davids, W.G. (1998). Modeling of Rigid Pavements: Joint Shear Transfer Mechanisms and Finite Element
Solution Strategies. PhD Dissertation, University of Washington, Seattle, WA.
Davids, W.G. and Mahoney, J. (1999). Experimental Verification of Rigid Pavement Joint Load Transfer
Modeling with EverFE. Transportation Research Record 1684, TRB, National Research Council,
Washington, D.C., pp. 81-89.
Davids, W.G. and Turkiyyah, G.M. (1999). Multigrid Preconditioner for Unstructured Nonlinear 3D FE
Models. Journal of Engineering Mechanics, ASCE, 125(2):186-196.
Davids, W.G. (2000). Effect of Dowel Looseness on Response of Jointed Concrete Pavements. Journal of
Transportation Engineering, ASCE, 126(1):50-57.
Davids, W.G., Wang, Z.M., Turkiyyah, G., Mahoney, J. and Bush, D. (2003). 3D Finite Element Analysis of
Jointed Plain Concrete Pavement with EverFE2.2. Transportation Research Record, TRB, National
Research Council, Washington, D.C. (in press).
Dei Poli, S., Di Prisco and Gambarova, P.G. (1992). Shear Response, Deformations, and Subgrade Stiffness
of a Dowel Bar Embedded in Concrete. ACI Structural Journal, 89(6):665-675.
Fish, J., Pan, L., Belsky, V. and Gomaa, S. (1996) Unstructured Multigrid Method for Shells. International
Journal for Numerical Methods in Engineering, 39:1181 1197.
Hjelmstat, K.D., Kim, J. and Zuo, K.H. (1997). Finite Element Procedures for Three-Dimensional Pavement
Analysis. Aircraft/Pavement Technology: In the Midst of Change, pp. 125-137, ASCE.
18
Ioannides, A.M. and Korovesis, G.T. (1990). Aggregate Interlock: A Pure Shear Load Transfer Mechanism.
Transportation Research Record 1286, TRB, National Research Council, Washington, D.C., 2001, pp. 14
24.
Ioannides, A.M. and Korovesis, G.T. (1992). Analysis and Design of Doweled Slab-on-Grade Pavement
Systems. Journal of Transportation Engineering, 118(6):745-768.
Jensen, E.A. and Hansen, W. (2003). A New Model for Predicting Aggregate Interlock Shear Transfer in
Jointed Concrete Pavements. Proceedings of EM2003-The 16th ASCE Engineering Mechanics
Conference, Seattle, WA, July 16-18, (CD-ROM).
Kuo, C., Hall, K. and Darter, M.. Three-Dimensional Finite Element Model for Analysis of Concrete
Pavement Support. Transportation Research Record 1505, TRB, National Research Council,
Washington, D.C., 1996, pp. 119 127.
Rasmussen, R.O. and Rozycki, D.K. (2001). Characterization and Modeling of Axial Slab-Support
Restraint. Transportation Research Record 1778, TRB, National Research Council, Washington, D.C.,
pp. 26 32.
Saad, Y. (1996). Iterative Methods for Sparse Linear Systems. PWS Publishing Co., Boston, MA.
Walraven, J.C. (1981). Fundamental Analysis of Aggregate Interlock. Journal of the Structural Division,
ASCE, 107(ST11):2245 2270.
Walraven, J.C. (1994). Rough Cracks Subjected to Earthquake Loading. Journal of Structural Engineering,
120(5):1510 1524.
Wang, C-K. and Salmon , C.G. (1985). Reinforced Concrete Design (4
th
Ed.). Harper and Row, New York.
Wattar, S.W., Hawkins, N.M. and Barenberg, E.J. (2001). Aggregate Interlock Behavior of Large Crack
Width Concrete Joints in PCC Airport Pavements. Technical Report, Department of Civil and
Environmental Engineering, University of Illinois at Urbana-Champaign.
Zhang, J. and Li, V.C. (2001). Influence of Supporting Base Characteristics on Shrinkage-Induced
Stresses in Concrete Pavements. Journal of Transportation Engineering, ASCE, 127(6);455 642.
Zienkiewicz, O.C. and Taylor, R.L. (1994). The Finite Element Method, Volume 1 (4
th
Ed.). McGraw Hill
Book Company, London.

S-ar putea să vă placă și