You are on page 1of 9

TRANSPORTATION RESEARCH RECORD 1684 Paper No.

99-0105 81

Experimental Verification of Rigid Pavement


Joint Load Transfer Modeling with EverFE
WILLIAM G. DAVIDS AND JOE P. MAHONEY

The joint load transfer modeling capabilities of EverFE, a recently lem of shear transfer across discrete cracks in concrete has shown
developed rigid pavement three-dimensional finite element analysis the mechanics of aggregate interlock shear transfer to involve a
tool, are verified through comparisons with available experimental complex interaction between two deformable, rough surfaces. In
data. Dowel joint load transfer is examined via comparison of dis-
addition to contact between the rough joint surfaces, there typically
placements predicted by EverFE with results from laboratory tests of
two small-scale doweled pavement systems, and dowel looseness is is localized crushing, the amount of which depends on the joint
shown to be a probable cause for experimentally observed differential opening, normal restraint of the joint, the strength of the concrete,
joint displacements. Results of finite element analyses using EverFE’s and the size and distribution of the aggregate particles (12). Despite
nonlinear, two-phase aggregate interlock constitutive model are shown these complexities, most finite element models of rigid pavements
to agree well with available experimental data. A parametric study is idealize aggregate interlock with linear springs spanning the joint
performed that examines the effect of joint opening on aggregate inter-
(2,13,14) or Coulomb friction with contact (1,3). Although the use
lock load transfer and illustrates the importance of considering non-
linearities in joint load transfer when predicting pavement response. of linear springs may be reasonable for an examination of the effect
Recommendations for future research on joint load transfer modeling of aggregate interlock shear transfer effectiveness on the global slab
are also discussed. response (13), the rational choice of a spring stiffness may be diffi-
cult, if not impossible. Coulomb friction, although playing an impor-
tant role in aggregate interlock load transfer, does not wholly account
The realistic modeling of both aggregate interlock and dowel load for the complexities inherent in aggregate interlock because of the
transfer across rigid pavement joints is necessary for the accurate idealization of the joint surfaces as smooth.
prediction of pavement response to applied axle and thermal load- EverFE, a recently developed program for the 3D finite element
ings using the finite element method. Unfortunately, the complexi- analysis of rigid pavements (15), incorporates joint load transfer
ties inherent in the mechanics of joint load transfer make achieving modeling strategies that consider important aspects of dowel and
this requirement difficult. Although nonlinear three-dimensional aggregate interlock load transfer not previously addressed. In par-
(3D) finite element modeling is rapidly becoming a standard ana- ticular, EverFE implements an embedded dowel element that per-
lytical tool in research settings (1–3), methods for modeling both mits the explicit and rigorous consideration of dowel looseness (16)
aggregate interlock and dowel joint load transfer still warrant further and a nonlinear aggregate interlock constitutive model employing a
investigation. two-phase idealization of the slab concrete that accounts for aggre-
Dowel-slab interaction plays an important role in dowel load gate size, joint opening, and concrete strength (12,17). Two studies
transfer; in particular, dowel looseness has been shown experi- verifying these joint load transfer models via comparisons with
mentally (4,5) and analytically (6,7 ) to have potentially detrimen- available laboratory data (11,18) are presented. The importance of
tal effects on joint load transfer. Most two-dimensional (2D) finite incorporating load transfer nonlinearities is demonstrated through a
element analyses account for dowel-slab interaction by incorpo- parametric study, and conclusions and recommendations for future
rating a Winkler foundation between the dowel and surrounding research are also addressed.
slab (8,9). More sophisticated models recently have been devel-
oped that account for the finite embedded length of the dowel (10);
however, these still rely on the common assumption of a Winkler
MODELING STRATEGIES
foundation sandwiched between the dowels and surrounding slab.
Although the Winkler foundation modulus can be reduced to account
Discretization of Slab and Base
for dowel looseness, determination of the proper foundation mod-
ulus can be accomplished only through backcalculation using
The finite element models employed in this study are similar to that
measured data. Further, a backcalculated modulus is effectively a
shown in Figure 1. The slabs and base are linearly elastic and are
secant stiffness, valid only for one loading, geometry, and set of
discretized with 20-noded quadratic hexahedra. A Winkler founda-
material properties.
tion is used to represent the subgrade below the bottommost base
Colley and Humphrey (11) demonstrated that aggregate interlock
layer. Loss of contact between the slab and the upper base layer is
load transfer effectiveness and endurance depend on many factors,
explicitly considered by using a nodal contact approach, illustrated
including slab thickness, load magnitude, foundation type, subgrade
in Figure 1 (15,19). The model generation capabilities and solution
modulus, and aggregate shape. Research into the more general prob-
efficiency of EverFE are central in permitting the simulation of such
systems with relative ease on desktop computers. In addition, EverFE
W. G. Davids, Department of Civil and Environmental Engineering,
University of Maine, 5711 Boardman Hall, Orono, ME 04469-5711. J. P.
permits the consideration of various wheel and axle load configu-
Mahoney, Department of Civil and Environmental Engineering, University rations as well as linear temperature gradients through the slab
of Washington, Box 352700, Seattle, WA 98195-2700. thickness (not considered in the present study).
82 Paper No. 99-0105 TRANSPORTATION RESEARCH RECORD 1684

nodal displacements of the dowel as a function of the nodal displace-


ments of the solid element it is embedded within, debonding condi-
tions, and any gaps that may exist. Consider a single (unembedded)
dowel element having a nodal displacement vector, Ud and stiffness
matrix Kd. The displacement vector Ud can be expanded to include the
displacement vector of the solid element the dowel lies within, Ue:

U de = [ ]
Ud
Ue
(1)

The new displacement vector for the embedded dowel element can
be transformed back to Ud as follows:

U d = TUde (2)

The transformation matrix, T, incorporates debonding and gap


information for each dowel node.
It follows from the principle of virtual work that the stiffness
matrix of the embedded dowel, Kde, can be computed as

Kde = T T K d T (3)

Because nonlinearities arising from nodal contact as well as


debonding information are encapsulated in the tangent stiffness
matrix, Kde, this formulation is particularly amenable to inclusion in
a general nonlinear solver. Further details of the element deriva-
tion and implementation may be found in the work by Davids and
Turkiyyah (16).
FIGURE 1 FE mesh of rigid pavement system and contact
modeling.
Aggregate Interlock Modeling
Dowel Modeling EverFE relies on a two-phase aggregate interlock model developed
by Walraven (12,17) to generate nonlinear aggregate interlock crack
The dowel modeling technique used in this study relies on an embed- constitutive relations. The crack is assumed to follow the aggregate
ded finite element formulation for the dowels having the following boundaries, and the aggregate particles bearing on the paste are
features: assumed to be at the point of slip [see Figure 3(a)]. Walraven’s
model also assumes that the aggregate particles are graded accord-
• The dowels may be located exactly without regard to meshing
ing to a Fuller distribution and the maximum particle diameter, dmax,
of the slabs, as shown in Figure 2. and volumetric percentage of aggregate, pk, are model parameters.
• The dowels may be debonded relative to the slabs.
Assuming an ultimate strength of the cement paste, σpu, a coeffi-
• Gaps between the dowels and surrounding slabs (dowel loose-
cient of friction between the paste and aggregate of µ, and computing
ness) may be explicitly modeled by using a nodal contact approach. the projected contact areas Ax and Ay by using the deformed geome-
try, the forces Fx and Fy required for equilibrium can be determined
Three-noded, 18 degree of freedom quadratic beam elements are for a single aggregate particle diameter–embedment combination by
used to represent the portions of the dowel embedded in the slabs, and using Equations 4 and 5 [see Figure 3(a)].
a conventional two-noded shear beam is used to span the joint. The
embedded formulation of the dowel element relies on expressing the σ = σ pu ( Ax − µ Ay ) ( 4)

τ = σ pu ( Ay + µ Ax) (5)

By using the probability of occurrence for a particular embedment–


diameter combination derived by Walraven (12), the likely forces on
all particles may be summed to give the total forces acting on a crack
plane of unit area. Typical crack shear stress-displacement relations
predicted by the model are shown in Figure 3(b), where each curve
corresponds to a specific joint opening; although only five curves are
shown, 30 are typically generated to give a more complete definition
of the stress-displacement relations. A zero-thickness, 16-noded qua-
dratic isoparametric interface element is used to incorporate crack
constitutive relations in the finite element models developed with
EverFE. Full details of the element formulation and implementation
FIGURE 2 Embedded dowel element. may be found elsewhere (7).
Davids and Mahoney Paper No. 99-0105 83

FIGURE 3 Aggregate interlock modeling and shear stress-displacement relations:


(a) aggregate particle distribution and particle bearing on paste; and (b) aggregate
interlock shear stress–displacement relations.

MODEL VALIDATION: DOWEL LOAD TRANSFER slipping plastic drinking straws over their ends. The model was cast
in a steel reaction box, and the ends of the slabs were clamped
This section presents a comparison of predicted pavement response between two angles to prevent displacement and increase their
with measured laboratory data (18). In particular, dowel looseness is effective length. For further details on all tests, see Hammons (18).
shown to be a potential cause for the observed differential vertical
joint displacements of a scale model jointed plain concrete pavement
system subjected to corner loading. Finite Element Model

Of the six tests completed by Hammons, tests LSM-2 and LSM-5


Experimental Study have been chosen for comparison with EverFE. They were selected
as bounding realistic conditions: LSM-2 had no treated base, with
The basic system tested by Hammons (18) consisted of two 51-mm- the slabs founded directly on the rubber pad; LSM-5 had a 38-mm-
thick concrete pavement slabs separated by a 1.6-mm-wide joint thick, debonded, monolithic cement-treated base, requiring that sep-
constructed with smooth surfaces to prevent aggregate interlock [as aration of the slabs and base be modeled. The finite element model
shown in Figure 4(a)]. To simulate typical construction, the 12 dow- developed for comparison with the test data closely matches exper-
els were purposely debonded in one slab by applying grease and imental conditions. The rubber reaction block is treated as a dense
84 Paper No. 99-0105 TRANSPORTATION RESEARCH RECORD 1684

Dowel Modeling

As will be seen, the tests showed significant variations in vertical dis-


placement across the joint near the load. If the dowels are assumed to
have no looseness, EverFE predicts very little relative vertical dis-
placement between the two slabs at the joint. However, it is probable
that there is some looseness between the dowels and the slabs because
of construction methods and tolerances. In the present finite element
models, a constant gap, γ, was assumed between the dowels and the
unloaded slab along their embedded length, whereas perfect contact
was maintained between the dowels and the loaded slab, as shown in
Figure 5. Four dowel elements were located near the joint to better
capture potential contact with two more elements used to discretize
the remaining embedded portions of the dowel, giving a total of
13 potential contact points for each dowel. It must be noted that
EverFE currently does not permit a constant dowel looseness profile
to be automatically generated, and some independent postprocessing
of the finite element input data was necessary.

Comparison with Experimental Data

Two different finite element analyses were run for comparison with
test LSM-2: the first assumed no dowel looseness; the second run
assumed a uniform gap, γ, of 0.08 mm between the dowel and the
unloaded slab. This value of the gap was not measured, but it was
backcalculated to give a relative vertical displacement between the
two loaded and unloaded slabs that closely matches the experimen-
tally measured values. To place this value in perspective, the mea-
FIGURE 4 Experimental model and finite element mesh:
(a) model dimensions and instrumentation layout; and
sured wall thickness of a typical drinking straw is approximately
(b) finite element mesh for LSM-5. 0.13 mm—nearly 60 percent greater than γ. The vertical displace-
ment of the top surface of the slabs was interpolated along the line
y = 432 mm to give the analytical deflection profile corresponding
to that measured experimentally, as shown in Figure 6(a).
liquid, and the slab and treated base are modeled as elastic con-
tinua. Figure 4(b) shows the finite element mesh used to model test
LSM-5; an identical model without the base layer was used for
LSM-2.
The material properties used in the finite element model were
those determined experimentally by Hammons. The slab and base
layer moduli, E, were taken as 27 600 MPa and 1410 MPa, respec-
tively; the corresponding values of Poisson’s ratio, ν, were 0.18
and 0.20. For the dowels, E = 200 000 MPa and ν = 0.30. The rub-
ber pad had a modulus, k, of 0.09 MPa/mm as backcalculated with
a 3D finite element model developed by Hammons to give good
deflection profile comparisons with tests of a single undoweled
slab founded directly on the rubber pad and subjected to corner
loading.
The model boundary conditions were chosen to reflect those of
the laboratory tests. The displacements measured at transducers D7
through D10 near the slab corners [see Figure 4(a)] were maintained
at the slab ends. Although the slabs were cast against the reaction
box walls (which were coated with a form release agent), the sides
of the slabs and base layers (± y faces) were not restrained. There
were three reasons for this: (a) the slabs were restrained only when
displacing in the +y direction on the +y face and in the −y direction
on the −y face; (b) even if contact between the slabs and box walls
occurred, the restraint provided by the box walls was not that of a
rigid support; and (c) shrinkage of the slabs could have resulted in a
gap between the slabs and box walls before loading, which had to
be overcome before any contact could occur. FIGURE 5 Assumed profile of gap around dowels.
Davids and Mahoney Paper No. 99-0105 85

placement at the joint is quite accurate, validating the 0.08-mm gap


determined from the analysis of LSM-2. Figure 6(b) shows signifi-
cantly better prediction of displacements when k = 0.07 MPa/mm is
assumed, in particular near the joint region. However, the maximum
error observed is about 17 percent, larger than that for LSM-2, and
the predicted slope of the deflection basin is significantly steeper
than that observed experimentally. Possible factors for these dis-
crepancies include inaccurate modeling of boundary conditions and
nonlinear and creep response of the rubber block not accounted for
in the finite element model.

MODEL VALIDATION: AGGREGATE INTERLOCK

To verify the applicability of EverFE’s aggregate interlock model, it


must be compared with available test data. Although controlled lab-
oratory tests on aggregate interlock in rigid pavements are scarce, the
work of Colley and Humphrey (11) does provide some validation of
the model as detailed in this section.

Study by Colley and Humphrey

Colley and Humphrey (11) performed a series of laboratory tests on


plain jointed concrete slabs 178 mm and 229 mm thick to quantify
the effectiveness and degradation of aggregate interlock load trans-
fer at transverse contraction joints [see Figure 7(a)]. The joints were
formed with a 25.4-mm-deep pour strip at the bottom of the joint and
a 25.4-mm-deep groove cut in the freshly poured concrete at the top
of the joint. The basic test rig consisted of a waterproof reinforced
concrete base box to hold the base and subgrade material, a frame
for restraining the slab ends from displacing vertically and thereby
increasing their effective length, and a device for adjusting and
maintaining joint opening, w. Repetitive loads were applied sequen-
tially to both sides of the joint via hydraulic rams acting on 406-mm-
diameter steel bearing plates to simulate a 40-kN wheel load crossing
the joint. The slabs had an average 28-day compressive strength of
38 MPa. In the tests used for comparison in the present study, the
base consisted of a 152-mm compacted gravel layer over a silty clay
FIGURE 6 Deflection basin comparisons for LSM-2 and subgrade; the measured composite modulus of subgrade reaction, k,
LSM-5 at y 5 432 mm.
was 0.039 MPa/mm.

If γ = 0, there is nearly perfect load transfer efficiency at the


joint, whereas γ = 0.08 mm gives an accurate relative vertical dis- Finite Element Models
placement at the joint as well as a better overall prediction of the
displacement basin. In general, the agreement between the finite The finite element models of the system tested by Colley and
element model and the experimental data is good, with a maximum Humphrey match the test conditions as closely as possible. The slabs
discrepancy of about 13 percent observed at x = 600 mm. were modeled as linearly elastic with a modulus of elasticity of 31 000
Two different finite element analyses were run for comparison MPa and Poisson’s ratio, ν, of 0.20. The base was treated as a dense
with test LSM-5, each with a fixed value of γ of 0.08 mm based on liquid with a measured k of 0.039 MPa/mm. The vertical displace-
the results of LSM-2. However, the first analysis was conducted with ments of the slab ends were fixed, and sufficient restraints were pro-
the previously backcalculated k for the rubber pad of 0.09 MPa/mm, vided to prevent rigid body displacements and rotations of the slabs.
and the second analysis assumed k = 0.07 MPa/mm. This reduction As shown in Figure 7(b), the finite element models employed four
is plausible: the value for k of 0.09 MPa was backcalculated by Ham- elements through the slab thickness, with a total of 960 quadratic
mons from comparison with a test of a single corner loaded slab rest- hexahedral elements and 15,666 degrees of freedom. Figure 7(c)
ing directly on a dense liquid. Lack of a treated base layer will increase gives details of the mesh near the joint, illustrating the local modi-
the stresses applied to the rubber pad; because it is a strain-hardening fications of the mesh required to prevent aggregate interlock load
material, the stiffness backcalculated from this test will be too high transfer over the upper and lower 25.4 mm of the joint.
for case LSM-5. The aggregate interlock constitutive relations were developed
Figure 6(b) compares the predicted and experimental deflection assuming σpu = 55 N/mm2, computed from the following relationship
basins. For both finite element models, the predicted relative dis- proposed by Walraven (17):
86 Paper No. 99-0105 TRANSPORTATION RESEARCH RECORD 1684

2 du
JTE = ( 7)
dl + du

where du is the vertical displacement of the unloaded slab and dl is


the vertical displacement of the loaded slab. A more commonly used
measure is displacement load transfer efficiency, LTE, which may
be computed from

du JTE
JTE = = (8)
dl 2 − JTE

Comparisons of the computed and experimentally measured LTE


values are given in Figure 8 for both the 178-mm and 229-mm slabs.
Figure 8 shows that the proposed two-phase model is capable of cap-
turing the observed variation in LTE with w. Although the values
are more accurate for the 178-mm slab, the overall trends are pre-
dicted reasonably well for both cases. It should be noted that greater

FIGURE 7 Details of tests and finite element model used for


verification: (a) plan of test slabs, (b) finite element mesh, and
(c) detail of mesh at joint.

σ pu = 8.0 fcc (6)

where fcc is the compressive strength of a 150-mm cube, taken as


1.25fc. The paste-particle coefficient of friction, µ, was fixed at 0.40, a
typical value recommended by Walraven (17). The maximum particle
diameter was set at the reported value of 38 mm (11), and pk = 0.75.
Sample curves for various joint openings are shown in Figure 3(b); a
total of 30 curves for values of w between 0.01 mm and 3.0 mm was
generated to define the constitutive relations at the joint.

Comparison of Experimental and Analytical Results

Colley and Humphrey reported values of joint efficiency, JTE, for FIGURE 8 Comparison of measured and computed LTE:
various values of joint opening, w, and number of load cycles. Joint (a) results for 178-mm-thick slab; and (b) results for 229-mm-
efficiency is defined according to thick slab.
Davids and Mahoney Paper No. 99-0105 87

accuracy may be achieved by backcalculating the parameters σpu ment was used for the upper layer of interface elements [Figure 9(c)].
and µ to fit these particular test results. It must be noted that although EverFE implements the two-phase
nonlinear aggregate interlock model, some manual modification of
the input files was necessary to include these zero-stiffness elements.
ANALYTICAL STUDY The 150-mm-thick base layer was assumed to be a compacted
gravel with E = 150 MPa and ν = 0.20. Below the base layer, 300 mm
This section illustrates the importance of modeling nonlinearities in of natural soil was modeled as a linearly elastic continuum having
joint shear transfer via a parametric study on aggregate interlock E = 75 MPa and ν = 0.20. A dense liquid was used to represent the
shear transfer. In particular, the effect of joint opening, w, on system remaining natural soil with a modulus of 0.054 MPa/mm. The elas-
response is examined. The importance of accounting for aggregate tic material properties of the slab were E = 28 000 MPa and ν = 0.25,
interlock nonlinearity is emphasized by comparing results of models and the mass density, ρ, was 2400 kg/m3. To develop the two-phase
assuming equivalent linear joint springs with those employing the aggregate interlock model, µ = 0.4 and σpu = 45 N/mm2 were taken as
two-phase model for varying magnitudes of axle loads. typical; pk was fixed at 0.75 and dmax set to 20 mm.

Model Details Effect of Joint Opening

The parametric studies employ a system similar to that shown previ- To study the effect of joint opening, a single, transversely centered
ously in Figure 1; a plan view of the model showing boundary con- 80-kN axle load was applied at the joint. As the joint opening, w,
ditions and loading is given in Figure 9(a). The slabs were 230 mm varies from 0.10 mm to 2 mm (a typical seasonal maximum for 3-
thick; for simplicity, the joint was not skewed, and the boundaries of to 5-m joint spacings), LTE decreases nearly 60 percent, as shown
the subgrade were not extended beyond the edges of the slabs. Axle in Figure 10(a).
loadings were applied at the joint through four wheels, each acting Contact conditions at the joint were monitored during the solu-
over a rectangular area 180 mm long by 180 mm wide (selected to tion, and the maximum relative normal displacement of the joint
give a realistic uniform pressure of approximately 0.62 MPa for an surfaces was approximately the minimum joint opening of 0.10 mm.
80-kN axle). The finite element mesh used in the analyses had approx-
Commensurate with the change in LTE is a rapid decrease in shear
imately 29,000 degrees of freedom and is shown in Figure 9(b). The
stresses on the joint surfaces with increasing w. The maximum shear
sawcut used at the joint was assumed to extend through one-third of
stress, corresponding to w = 0.10 mm, is 204 kPa; this is reduced
the pavement thickness; to reflect this detail, a zero-stiffness joint ele-
more than 50 percent for w = 0.40 mm and to only 11 kPa for
w = 1.60 mm. The maximum principal stresses occur at the bottom
of the loaded slab under the wheel loads and range from 704 kPa for
w = 0.10 mm to 1063 kPa for w = 2.0 mm, an increase of 51 percent;
as w increases, the loading approaches that of an unprotected edge.

Significance of Nonlinear Load Transfer

A separate group of analyses was also performed assuming values


of linear joint stiffnesses, q, ranging from 0 to 5 MPa/mm; LTE as
a function of q is also shown in Figure 10(a). While Figure 10(a) can
be used to compute an equivalent q for a given joint opening, this
linear stiffness is valid only for one geometry, loading, and set of
material properties. To illustrate, consider the case in which
LTE = 80 percent for an 80-kN axle load: from Figure 10(a), q = 0.5
MPa/mm and w = 0.6 mm. A separate group of finite element
analyses run for the same system with w fixed at 0.6 mm and var-
ious load levels predicts increasing LTE with increasing axle load,
as shown in Figure 10(b). This may be explained by the slab kine-
matics: at higher load levels, the joint opening decreases at the top
of the slabs, increasing the effectiveness of the aggregate interlock
load transfer. Conversely, the model with a fixed value of q = 0.5
MPa/mm predicts a decline in load transfer with increasing axle
load, as shown in Figure 10(b).
The implication of this analysis is clear: a backcalculated linear
spring stiffness should be used with caution when different loadings
are considered. As a practical case, if a linear spring stiffness is back-
calculated by using data from falling-weight deflectometer readings
or other nondestructive testing methods, analyses employing this
FIGURE 9 Model used for parametric studies: (a) plan of value of q are strictly valid only for the load level used in the non-
model and applied axle loading; and (b) finite element mesh. destructive test. Although only aggregate interlock was considered
88 Paper No. 99-0105 TRANSPORTATION RESEARCH RECORD 1684

FIGURE 10 Effect of joint opening, linear stiffness, and load level on


LTE: (a) LTE as a function of linear joint stiffness and joint opening;
and (b) LTE versus load level for q = 0.5 MPa/mm and w = 0.6 mm.

in the present study, nonlinearities due to dowel looseness will study on the effect of joint load opening on aggregate interlock load
present similar difficulties. transfer was conducted.
In general, the validation studies indicated that the aggregate
interlock and dowel load transfer modeling capabilities of EverFE
CONCLUSIONS are reasonably accurate and viable. Further, the parametric study
results showed that joint load transfer nonlinearities can have sig-
This study presented experimental verification of the nonlinear joint nificant effects on pavement response, and a backcalculated linear
load transfer modeling strategies employed by EverFE. Dowel loose- spring stiffness should be used with caution when different loadings
ness was shown to be a probable cause for observed differential ver- are considered.
tical joint displacements in the scale model rigid pavement tests of Although the modeling techniques verified here represent sig-
Hammons (18), and results predicted by the two-phase aggregate nificant advances in quantifying aggregate interlock and dowel
interlock model compared reasonably well with the data of Colley joint load transfer, several issues should be addressed in future
and Humphrey (11). In addition to model validation, a parametric research:
Davids and Mahoney Paper No. 99-0105 89

• Both the loss of aggregate interlock shear transfer efficiency TRB, National Research Council, Washington, D.C., 1994, pp.
and the formation of dowel looseness under cyclic loading should 123–133.
4. Snyder, M. B. Cyclic Shear Load Testing of Dowels in PCC Pavement
be quantified. This likely would require significant extension to and Repairs. In Transportation Research Record 1215, TRB, National
refinement of the existing capabilities of EverFE. Research Council, Washington, D.C., 1989, pp. 246–257.
• More controlled laboratory studies under the high-cycle, low- 5. Buch, N., and D. G. Zollinger. Development of Dowel Looseness Pre-
stress conditions common to pavements should be conducted to diction Model for Jointed Concrete Pavements. In Transportation
Research Record 1525, TRB, National Research Council, Washington,
permit construction and calibration of more refined models. D.C., 1996, pp. 21–27.
• Field verification is necessary to provide a higher degree of 6. Zaman, M., and A. Alvappillai. Contact-Element Model for Dynamic
confidence in the model predictions. Analysis of Jointed Concrete Pavements. Journal of Transportation
Engineering, Vol. 121, No. 5, 1995, pp. 425– 433.
7. Davids, W. G., G. M. Turkiyyah, and J. Mahoney. Modeling of Rigid
Achieving these future research objectives would lead to more Pavements: Joint Shear Transfer Mechanisms and Finite Element Solu-
reliable and accurate quantification of aggregate interlock and tion Strategies. Report WA-RD 455.1. Washington State Department of
dowel load transfer. It must also be noted that the effects of tem- Transportation, Olympia, 1998.
8. Tabatabaie, A. M., and E. J. Barenberg. Structural Analysis of Concrete
perature and moisture-induced slab curling should be considered in Pavements. Journal of the Transportation Division, ASCE, Vol. 106,
future studies; EverFE currently permits the inclusion of a linear tem- No. TE5, 1980, pp. 832–849.
perature gradient. Integration of such models with a reliable, effi- 9. Tayabji, S. D., and B. E. Colley. Analysis of Jointed Concrete Pave-
cient, and user-friendly 3D finite element package such as EverFE ments. Report FHWS-RD-86-041. FHWA, U.S. Department of Trans-
portation, 1986.
and their use in conjunction with appropriate nondestructive joint 10. Guo, H., J. A. Sherwood, and M. B. Snyder. Component Dowel-Bar
evaluation techniques would aid researchers, designers, and planners Model for Load-Transfer Systems in PCC Pavements. Journal of
in determining pavement retrofit and maintenance schedules. Transportation Engineering, Vol. 121, No. 3, pp. 289–298, 1995.
11. Colley, B. E., and H. M. Humphrey. Aggregate Interlock at Joints in
Concrete Pavements. Bulletin 189, HRB, National Research Council,
Washington, D.C., 1967, pp. 1–18.
ACKNOWLEDGMENTS 12. Walraven, J. C. Fundamental Analysis of Aggregate Interlock. Journal of
the Structural Division, ASCE, Vol. 107, No. ST11, 1981, pp. 2245–2270.
13. Ioannides, A. M., and G. T. Korovesis. Aggregate Interlock: A Pure-
This work was supported in part by WSDOT under contract T9903-54 Shear Load Transfer Mechanism. In Transportation Research Record
and by fellowships from the Valle Program and the Osberg Foun- 1286, TRB, National Research Council, Washington, D.C., 1990, pp.
dation at the University of Washington. Michael Hammons of Applied 14–24.
14. Brill, D. R., G. F. Hayhoe, and X. Lee. Three-Dimensional Finite Ele-
Research Associates, Inc., is gratefully acknowledged for providing ment Modeling of Rigid Pavement Structures. In Aircraft/Pavement
the doweled slab laboratory data referenced in this manuscript. Technology: In the Midst of Change, ASCE, 1997, pp. 151–165.
15. Davids, W., G. Turkiyyah, and J. Mahoney. EverFE—Rigid Pavement
Three-Dimensional Finite Element Analysis Tool. In Transportation
Research Record 1629, TRB, National Research Council, Washington,
REFERENCES D.C., 1998, pp. 41–49.
16. Davids,W., and G. Turkiyyah. Development of Embedded Bending
Member to Model Dowel Action. Journal of Structural Engineering,
1. Uddin, W., R. M. Hackett, A. Joseph, Z. Pan, and A. B. Crowley. Three- Vol. 123, No. 10, 1997, pp. 1312–1320.
Dimensional Finite-Element Analysis of Jointed Concrete Pavement 17. Walraven, J. C. Rough Cracks Subjected to Earthquake Loading. Jour-
with Discontinuities. In Transportation Research Record 1482, TRB, nal of Structural Engineering, Vol. 120, No. 5, 1994, pp. 1510–1524.
National Research Council, Washington, D.C., 1995, pp. 26–32. 18. Hammons, M. I. Development of an Analysis System for Discontinuities
2. Kuo, C.-M., K. T. Hall, and M. I. Darter. Three-Dimensional Finite Ele- in Rigid Airfield Pavements. Report GL-97-3. U.S. Army Corps of Engi-
ment Model for Analysis of Concrete Pavement Support. In Trans- neers, 1997.
portation Research Record 1505, TRB, National Research Council, 19. Davids, W., and G. Turkiyyah. Multigrid Preconditioner for Unstruc-
Washington, D.C., 1995, pp. 119–127. tured Nonlinear 3D FE Models. Journal of Engineering Mechanics,
3. Zaghloul, S. M., T. D. White, and T. Kuczek. Evaluation of Heavy Load Vol. 125, No. 2, 1999, pp. 186–196.
Damage Effect on Concrete Pavements Using Three-Dimensional, Non-
linear Dynamic Analysis. In Transportation Research Record 1449, Publication of this paper sponsored by Committee on Rigid Pavement Design.

You might also like