Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

CHAPTER 2

LITERATURE REVIEW

This chapter presents a survey of existing literature on theories and research

techniques related to concrete pavement slab analysis, covering studies on structural

analysis of concrete pavements under the action of vertical loads, and studies on the effect

of temperature distribution on warping stresses in concrete pavements.

2.1 Structural Analysis of Concrete Pavements under Vertical Loads

This section provides a review on foundation models adopted for the problems of

structure-foundation interaction, and analysis models for concrete pavement response

under traffic loading.

2.1.1 Foundation Models

The analysis of slabs-on-grade type pavements (without base layer) is a form of soil-

structure interaction analysis problems of interest in the fields of highway structure and

geotechnical engineering. As in numerous other engineering applications, the response of

the supporting soil medium under the pavement is an important consideration. To obtain

an accurate evaluation of this response, it is necessary to capture the complete stress-strain

characteristics of the foundation. Accurately describing the stress-strain characteristics of

any given foundation medium is usually hindered by the complex soil conditions, which

are markedly nonlinear, irreversible and time dependent. Furthermore, these soils are

generally anisotropic and inhomogeneous. Idealized models have been developed to

7
simulate soil response under predefined loading and boundary conditions. Certain

assumptions about the soil medium are used for these idealizations. They are necessary for

reducing the analytical rigor of such a complex boundary value problem. Two of the most

frequently applied assumptions are that of linear elasticity and homogeneity.

2.1.1.1 Dense Liquid Foundation Model

In the dense liquid foundation model, also known as the Winkler (1867) foundation

model, the foundation is considered as a bed of evenly spaced, independent, linear springs.

The model assumes that each spring deforms in response to the vertical stress applied

directly to that spring, and does not transmit any shear stress to the adjacent springs. The

relation between an external load p applied on any point of the surface of the foundation

and the deflection w of foundation at that point is given by the Equation:

p = kw (2-1)

where k is the modulus of subgrade reaction.

No transmission of shear forces means that there are no deflections beyond the edges

of the plate or slab. The liquid idealization of this foundation type (illustrated in Figure 2.1)

was derived for its behavioral similarity to a medium following Archimedes’ buoyancy

principle – the weight of a boat is equal to the water displaced. It was applied to analyze

pavement support systems in studies by Westergaard (1926b, 1933, and 1947).

Figure 2.1 Dense Liquid and Elastic Solid Extremes of Elastic Soil Response

8
In the field, the k-value is determined using data obtained from a plate loading test

performed on the foundation using a 30-inch (0.76m) diameter plate (Ioannides et al.

1985). The load is applied to a stack of 1-inch (0.0245m) thick plates, until a specified

pressure (p) or deflection (Δ) is reached. The k-value is then computed as the ratio of the

pressure to the corresponding deflection, i.e.,

p
k= (2-2)
Δ
Another method for obtaining a k-value for use in analysis is by backcalculation from

measured deflections of the slab surface obtained from non-destructive tests using devices

such as the Falling Weight Deflectometer (FWD).

2.1.1.2 Elastic Solid Foundation

The elastic solid foundation model, sometimes referred to as the Boussinesq

foundation, treats the soil as a linearly elastic, isotropic, homogenous material that extends

semi-infinitely. It is considered to be a more realistic model of subgrade behavior than the

dense liquid model because it can take into account of the effects of shear transmission of

stresses to adjacent support elements. Consequently, the distribution of displacements are

continuous., the deflection of a point in the subgrade occurs not just as a result of the

stress acting at that particular point, but is also influenced to a progressively decreasing

extnt by stresses at points further away.

The analysis of loaded slab supported on a solid foundation is mathematically a much

more difficult problem for which a number of solutions were available in the literature,

such as the Boussinesq solution given below

2(1 − μ s2 ) pr0
w=
Es (2-3)

9
where w = displacement of foundation surface at the center of loaded area, p =

contact pressure, r0 = radius of the loaded area, μ s= Poisson’s ratio of foundation, and Es=

elastic modulus of foundation.

Due to its mathematical complexity, the solid foundation model has been less

attractive than the dense liquid foundation model. Unlike the dense liquid foundation

model, where the governing Equations are of a differential form, the elastic foundation

model requires the solution of integral or integro-differential Equations. The continuous

nature of the displacement function in the elastic solid model also means that this model

cannot accurately simulate pavement behavior with discontinuities in the structure,

especially for slabs supported on natural soil subgrades. The model is unsuitable for

predicting slab responses at edges, corners, cracks or joints with no physical load transfer.

The elastic solid foundation model considers the shear force interaction between

different elements in the foundation. Although it presents an improvement over Winkler

foundation model by considering the shear forces in the foundation, field tests showed that

the solutions were not exact for many foundation materials. Foppl (1909) reported that the

surface displacements of foundation soils outside the loaded region decreased faster than

the prediction by this model.

2.1.1.3 Improved Foundation Models by Modification of Winkler Foundation

The dense liquid and elastic solid foundation models may be considered as two

extreme idealizations of actual soil behavior. The dense liquid model assumes complete

discontinuity in the subgrade and is better suited for soils with relatively low shear

strengths (e.g. natural subgrade soils). In contrast, the elastic solid model emulates a

perfectly continuous medium and is better suited for soils with high shear strengths (e.g.,

treated bases). The elastic response of a real soil subgrade lies somewhere between these

10
two extreme foundation models. In real soils, the displacement distribution is not

continuous, neither is it fully discontinuous. The deflection under a load can occur beyond

the edge of the slab and it goes to zero at some finite distance. In an attempt to bridge the

gap between the dense liquid and elastic solid foundation models, researchers have

developed some improved foundation models. These improved foundation models were

developed in either of the following two ways: (a) starting with the Winkler foundation

and, in order to bring it closer to reality, some kind of interaction between spring elements

may be assumed, or (b) starting with the elastic solid foundation, simplifying assumptions

with respect to expected displacements or stresses may be introduced.

A major problem in applying these models, however, has been the lack of guidance

in selecting the governing parameters which have limited or no physical meaning. Brief

overview of some improved foundation models are given below.

Filonenko – Borodich Foundation Model

The Filonenko-Borodich foundation model is perhaps one of the earliest two-

parameter models. In addition to the vertical springs used to simulate the dense liquid

foundation model, this foundation model includes a stretched elastic membrane that

connects to the top of the springs and is subjected to a constant tension field T. The

tension membrane allows for interaction between adjacent spring elements. The relation

between the subgrade surface stress field p and the corresponding deflection w is defined

by

p = kw − T∇ 2 w (2-4)
where ∇ 2 = the Laplace operator and T = constant tension. A schematic of the Filonenko-

Borodich model is given in Figure 2.2.

11
Figure 2.2 Filonenko-Borodich Foundation Model
Hetenyi Foundatio

Hetenyi (1946, 1950) proposed that the interaction between the independent spring

elements be accomplished by embedding in the two-dimensional case an elastic beam and

in the three-dimensional case a plate in the material of the Winkler foundation. It is

assumed that the beam or plate deforms in bending only. The relation between contact

pressure p and deflection of foundation surface w for the three-dimensional case is

p = kw + D s ∇ 2 ∇ 2 w
(2-5)

where =2 the Laplace operator, and = the flexural rigidity of an imaginary plate in the
∇ Ds
Winkler foundation to represent interaction between independent spring elements.

Pasternak Foundation

Pasternak (1954) considered the shear interactions between the spring elements of

Winkler foundation by connecting the ends of the springs with a beam or plate consisting

of incompressible vertical elements which deformed only by transverse shear. Under this

assumption, the relation between the contact pressure p and deflection of foundation

surface w is given by

p = kw + Gb ∇ 2 w
(2-6)

where = shear modulus of foundation. A schematic of the Pasternak model is given in


Gb
Figure 2.3.

12
Shear Layer (Gb)_

Figure 2.3 Pasternak Foundation Model

“Genelized” Foundation by Venckovskii

In this foundation model, in addition to the Winkler hypothesis, Venckovskii (1958)

assumed that the applied moment is proportional to the angle of rotation. Analytically
Mn
this is described by

p = kw (2-7a)
dw
M n = k1 (2-7b)
dn
where n is any direction at the point in the plane of the foundation surface, and k and k1
are the corresponding proportionality factors.
2.1.1.4 Improved Foundation Models by Modification of Elastic Solid Foundation

Reissner Foundation

Assuming that the in-plane stresses throughout the foundation layer are negligibly

small, i.e.,

σ x = σ y = τ xy = 0
(2-8)

and that the horizontal displacements at the upper and lower surfaces of the foundation

layer are zero, Reissner (1958) obtained for the elastic case the following relation

c2 2
c1 w − c 2 ∇ 2 w = p − ∇ p (2-9)
4c1
where
Es HG s Es
c1 = , c2 = , Gs = (2-10)
H 3 2(1 + μ )

13
To apply the Reissner model to the case in which elastic modulus E s varies linearly

with the depth of foundation, Horvath (1983) developed a modified Reissner model as

follows,

C1 w − C 2 ∇ 2 w = p − C 3 ∇ 2 p (2-11)

where C1 , C 2 and C 3 are constants which are functions of elastic modulus Eb and
thickness H of the foundation.
Beam-column-analogy Foundation

From an elastic continuum, Horvath (1989) developed a Pasternak-type, beam-

column-analogy foundation model as

Es GH
p= w − s ∇2w (2-12)
H 2

With this model, Harvath (1992) analyzed the mat-supported Chemistry Building at

Massachusetts Institute of Technology in Cambridge, Massachusetts. The comparison of

calculated and observed settlements showed that this model provided good agreement

with the observed behavior.

2.1.2 Analytical Solutions for Concrete Pavement Response to Traffic Loading

2.1.2.1 Goldbeck “Corner Formula”

The first attempt to rigid pavement mechanistic design and analysis as recorded in

literature was an approach by Goldbeck (1919) who proposed the “corner formula” for

stresses in concrete slab. This formula was based on the assumption that under a

concentrated load, the slab corner acted as a cantilever beam of variable width, receiving

no support from the subgrade between the corner and the point of maximum moment in

the slab. The tensile stress in the top of the slab may be computed as:

14
3P
σc = (2-13)
h2

where is the stress due to the corner load P and h is the thickness of the slab. A major
σc
disadvantage of this solution is that it neglects supports of concrete pavements from

subgrade, and so this solution greatly overestimates maximum bending stresses in

concrete pavements.

2.1.2.2 Westergaard’s Closed-Form Solutions

Westergaard (1926b, 1926c, 1927, 1933, 1939, 1943, 1948) proposed the first

complete theory of structural behavior of rigid pavements. Westergaard modeled the

pavement structure as a homogenous, isotropic, elastic, thin slab resting on a Winkler

(dense liquid) foundation. From existing test data and experience, he identified the three

most critical loading positions, the interior (also called center), edge, and corner, as

illustrated in Figure 2.4 and developed Equations for computing critical stresses and

deflections for loads placed at the edge, corner and center respectively.

Figure 2.4 Problem Definition and Three Loading Positions by Westergaard

Westergaard made the following simplifying assumptions in his analysis,

1. The concrete pavement acts as single semi-infinitely large, homogenous, isotropic

elastic slab with no discontinuities;

2. The foundation acts like a bed of springs (dense liquid foundation model) under the slab;

3. There is full contact between the slab and foundation;

15
4. All forces act normal to the surface where shear and frictional forces are negligible;

5. The semi-infinite foundation has no rigid bottom;

6. The slab is of uniform thickness, and the neutral axis is at its mid-depth.

7. The load at the interior and at the corner of the slab respectively is distributed uniformly

over a circular contact area; for the corner loading, the circumference of this circular area

is tangential to the edge of the slab.

8. The load at the interior edge of the slab is distributed uniformly over a semicircular

contact area, the diameter of the semicircle being along the edge of the slab.

In spite of limitations associated with the simplifying assumptions, Westergaard’s

Equations are still widely used today for computing stresses in pavements and validating

other models developed using different techniques.

Westergaard’s original Equations have been modified several times by different

authors, mainly to bring them into better agreement with measured responses of actual

pavement slabs. Ioannides et al (1985) performed a thorough study on Westergaard’s

original Equations and the modified formulas. They also compared the results with the

ILLI-SLAB finite element program. This comparison led to the development of new

Equations for the corner loading case.

Extensive investigations on the structural behavior of concrete pavement slabs

performed at the Arlington Experimental Farm (Teller and Sutherland 1943) showed

basically good agreement between observed stresses and those computed by Westergaard

theory, as long as the slab remained in full contact with the foundation. Proper selection of

the modulus of subgrade reaction was found to be essential for good agreement.

Interior Loading

Westergaard defines interior loading as the case when the load is at a “considerable

16
distance from the edge”. For this case the maximum bending stress at the bottom of the

slab due to the circular load of radius r0 is given by:

3(1 + μ ) P ⎛ L ⎞
σ= ⎜ ln + 0.6159 ⎟⎟ (2-14)
2πh 2 ⎜⎝ rc ⎠

where P = uniformly distributed circular pressure, h = slab thickness, E = elastic modulus

of concrete, μ = Poisson’s ratio of concrete, k = modulus of subgrade reaction, L is the

radius of relative stiffness defined as

1
⎡ Eh 3 ⎤4
L=⎢
( ⎥
⎣12 1 − μ k ⎦
2
) (2-15)

and,

rc = r0 when r0 ≥ 1.724h (2-16a)

rc = 1.6r02 + h 2 − 0.675h when r0 < 1.724h (2-16b)

The modified radius rc was introduced to account for the effect of shear stresses in the

vicinity of the load, which is neglected in the classical thin-plate theory. The deflection

Equation due to interior loading is given by (Westergaard 1939):

P ⎧⎪ 1 ⎡ ⎛ r0 ⎞ ⎤⎛ r0 ⎞
2
⎫⎪
w= ⎨1 + ⎢ln⎜ 2 L ⎟ − 0.673⎥⎜ L ⎟ ⎬ (2-17)
8kL2 ⎪⎩ 2π ⎣ ⎝ ⎠ ⎦⎝ ⎠ ⎪⎭

Corner Loading

Using a method of successive approximation, Westergaard proposed the following

formulas for computing the maximum bending stress and deflection, respectively, when

the slab is subjected to corner loading:

17
3P ⎡ ⎛⎜ r0 2 ⎞⎟ ⎤
0.6

σ = 2 ⎢1 − ⎜ ⎥ (2-18a)
h ⎢ ⎝ L ⎟⎠ ⎥
⎣ ⎦

P ⎡ r0 2 ⎤
w= ⎢1.1 − 0.88 ⎥ (2-18b)
kL2 ⎢⎣ L ⎥⎦

Edge Loading

Westergaard (1926b, 1933, 1948) defined edge loading as the case when “the wheel

is at the edge of the slab, but at a considerable distance from any corner”. Two possible

scenarios exist for this loading case: (1) a circular load with its center placed a radius

length from the edge, and (2) a semi-circular load with its straight edge in line with the

slab. The following Equations by Ioannides et al (1985) include modifications made to the

original Westergaard Equations. For the case of circular loading, the maximum bending

stress and deflection are computed as,

3(1 + μ ) P ⎡ ⎛ Eh 3 ⎞ 4 μ 1 − μ 1.18(1 + 2 μ )r0 ⎤


σ = ln⎜
2 ⎢ ⎜
⎟ + 1.84 −
2 ⎟
+ + ⎥ (2-19a)
π (3 + μ )h ⎣⎢ ⎝ 100kr0 ⎠ 3 2 2L ⎦⎥

2 + 1.2 μP ⎡ (0.76 + 0.4 μ )r0 ⎤


w= ⎢1 − ⎥ (2-19b)
Eh 3 k ⎣ L ⎦

The maximum bending stress and deflection for a semi-circular loading at the edge is,
3(1 + μ ) P ⎡ ⎛ Eh 3 ⎞ 4 μ (1 + 2 μ )r0 ⎤
σ = ln⎜
2 ⎢ ⎜
⎟ + 3.84 −
2 ⎟
+ ⎥ (2-20a)
π (3 + μ )h ⎣⎢ ⎝ 100kr0 ⎠ 3 2 L ⎦⎥

2 + 1.2 μP ⎡ (0.323 + 0.17 μ )r0 ⎤


w= ⎢1 − ⎥ (2-20b)
Eh 3 k ⎣ L ⎦

Due to simplifications associated with the assumptions stated above, several

limitations exist in the Westergaard theory. Some of these limitations are:

1. Stresses and deflections can be calculated only for interior, edge and corner loading

18
conditions;

2. Shear and frictional forces on the slab surface are ignored;

3. The Winkler foundation extends only to the edge of the slab. In reality, additional

support is provided by the surrounding subbase and subgrade;

4. The theory does not account for unsupported areas of the slab that results from voids or

discontinuities;

5. Multiple wheel loads cannot be considered; and,

6. Load transfer between joints or cracks is not considered when calculating the stresses or

deflections.

2.1.2.3 Other Solutions Based on Thin-Plate Theory

The classical thin-plate based theoretical models for structural analysis of concrete

pavement did not develop much further since Westergaard published his results. Although

influence chart developed by Pickett and Ray (1951) made Westergaard’s solution easy to

use and popular for the design of concrete pavements, further developments have received

less attention than they merit because of the complexities of the mathematics involved. It

is challenging to consider the finite dimensions of pavement slabs and the effect of joints

in analytical model due to the difficulty of expressing and solving the practical problems

mathematically.

Hogg and Hall (1938) took the subgrade as a semi-infinite elastic solid, and they

developed an analytical model for determining the stresses and deflections of a concrete

slab under the action of a single load by using the elastic properties of subgrade. This

model is effectively an infinite thin slab model because the derivation considers a single

interior load far away from any edge or corner of a slab.

Pickett and Ray (1951) further developed the work of Westergaard and Hogg to

19
include any arbitrary loading configuration, and developed a set of influence chats for

design. Their solutions are widely used for the application of Westergaard’s theory in

pavement design today.

2.1.2.4 Solutions Based on Thick-Plate Theory

The analytical solutions described in the preceding section are all based on thin- plate

theory that was first proposed by Kirchhoff in 1850. Reissner (1945, 1950 and 1958)

developed a thick-plate theory to analyze two problems: (1) the problem of torsion of a

rectangular plate, and (2) the problems of plain bending and pure twisting of an infinite

plate with a circular hole. The Reissner theory is based on modeling the plate structure as

two-dimensional structure with assumed stress variation through the plate thickness (the

third dimension). This theory takes into account shear deformation and transverse normal

stresses. The Reissner theory is regarded as stress-based shear deformable theory as it is

based on assumed stress variation through the plate thickness. Hu (1981) further extended

Reissner’s theory and developed another set of basic Equations for thick plates that are

simpler to solve than the original Equations.

Mindlin (1951) proposed another formulation to account for shear deformation based

on a proposed displacement field through the plate thickness. Mindlin’s theory is regarded

as a displacement-based shear deformable plate theory. Mindlin’s theory is also called the

first-order shear theory as it proposes a linear variation of the displacements through the

plate thickness. Other plate theories have been proposed which consider higher variations

of the displacement through the thickness.

A theoretical solution to the problem of a rectangular thick plate with four free edges

and supported on Pasternak foundation was developed by Shi et al. (1994). The

fundamental Equations for the problem were established by applying Reissner thick- plate

20
theory and solved by applying the method of superposition. Fwa et al (1996a) further

extended this solution into the analysis of concrete pavement and found differences

existed in both stresses and deflections between the thick-plate solutions and

Westergaard’s solution.

2.1.2.5 Solutions Based on Multiple-Slab Thick-Plate Models

To incorporate the effects of joint load transfer into the analysis of jointed concrete

pavement system, the researchers at the Transportation Research Center of the National

University of Singapore (NUS) have developed a three-slab thick-plate model (Shi 1995,

1996) and a five-slab thick plate model (Zhang 2001) that are able to predict the stresses

and deflections of concrete pavement systems under vertically applied wheel loads. The

three-slab model can deal with joints in one direction only. The five-slab model can

consider the effects of transverse and longitudinal joints at the same time. The three-

slab model is the first theoretical multiple slab model that incorporates the effects of joint

load transfer into the analysis of rigid pavements. This model was ideal for the analysis of

rigid pavement systems with only one lane. It cannot analyze the effects of slabs in the

adjacent lane(s) in a rigid pavement system with two or more lanes were analyzed using

this model. This can be treated as a limitation of the three-slab model. The five-slab model

was developed to consider the effects of joint load transfer in both longitudinal and

transverse directions. This is an improvement over the three-slab model and it can be used

to analyze rigid pavement systems with three or more lanes. However, the effects of four

corner slabs were ignored in the five-slab model and also there is no such five-slab system

in real rigid pavements.

2.1.3 Numerical Solutions for Concrete Pavement Response to Traffic Loading

It has been virtually impossible to obtain analytical closed-form solutions for many

21
pavement structures because of complexities associated with geometry, boundary

conditions, and material properties. The evolution of high-speed computers has facilitated

the analysis of such problem. The sections that follow are intended to provide a brief

background of some of the most commonly used numerical techniques for analyzing

concrete pavement structures.

2.1.3.1 The Discrete Element Method (DEM)

The first use of a discrete-element model for concrete pavement analysis was made

by Hudson and Matlock (1966). In this analysis the subgrade was idealized as a Winkler

foundation. The effects of joints in this model were taken into consideration by reducing

the original bending stiffness of the slab at those locations where a joint existed. The

model developed by Hudson and Matlock (1966) was later modified and improved by

Pearre and Hudson (1969) and Vora and Matlock (1970) to include elements of different

sizes, anisotropic skew slabs, the idealization of the subgrade as a semi-infinite elastic

solid. The major disadvantages of discrete-element formulations are that elements of

varying sizes are not easily incorporated into the analysis, and that special treatment is

needed at the free edge where stresses cannot be determined uniquely.

2.1.3.2 The Finite Element Method (FEM)

The finite-element method (FEM) is by far the most widely applied numerical

technique for the analysis of concrete pavements. It provides a modeling alternative that is

well suited for applications involving systems with irregular geometry, unusual boundary

conditions or non-homogenous composition. Finite-element models developed for

analysis of the concrete pavement system may be grouped into the following major classes:

(1) two-dimensional models, (2) slab models, and (3) three-dimensional models.

Two-dimensional models

22
Two-dimensional models consist of two types of models, plane strain models and

prismatic models. In the plane-strain idealization of the concrete pavement system,

changes in load or material along the longitudinal direction of the pavement are neglected.

The pavement system is represented as a transverse slice of the pavement having a unit

thickness. Because of their simplifying assumptions, these models are not capable of

evaluating various concrete pavement features such as the slab action of concrete

pavement, and the effects of multiple loads.

In the prismatic idealization of the concrete pavement system, the pavement system

is represented by a constant two-dimensional geometric shape with respect to an infinite

third dimension. The first prismatic model was developed by Wilson (1970). Wilson’s

model was later extended by Pichumani (1970), Crawford Pichumani (1975), and Tia et al.

(1987a, 1987b, 1988, 1989) for analysis of pavement structures. The main limitation of

these models for analysis of the concrete pavement is that no variation of geometrical

configuration is allowed along the longitudinal axis of the pavement system. Therefore,

prismatic models cannot handle any transverse discontinuities, nor do they use a realistic

representation for applied loads.

Slab models

The slab models currently available idealize concrete pavement slabs as classical thin

plates supported by an idealized subgrade, either the Winkler foundation or the semi-

infinite elastic solid foundation. Examples of slab models include KENSLAB developed

by Huang (1974, 1985, 1993), ILLI-SLAB (Tabatabaie 1978), WESLIQUID and

WESLAYER (Chou and Huang 1979, 1981), and FEACONS (Tia et al 1987, 1988).

(1) KENSLAB

In Huang’s computer model KENSLABS (Huang 1973, 1974, 1982, 1985, 1993), a

23
concrete pavement system with joints is represented by two or four slabs when load

transfer effects are considered, or one slab when no load transfer at joints is assumed. The

slab is treated in this model as composed of two bonded or unbonded layers with uniform

thickness. The two layers can be either a high modulus asphalt layer on top of a concrete

slab, a concrete slab, or a cement-treated base. Rectangular thin-plate elements with three

degrees of freedom per node (a vertical deflection and two rotations) are used to represent

the slab. Load transfer through doweled joint or aggregate interlock can be considered in

this model. For these two cases, the stiffness of joint is represented by a shear spring

constant CW and a moment spring constant Cθ defined as follows,

Shear force per unit length of jo int


Cw = (2-21a)
Difference in deflections between two slab

Moment per unit length of jo int


Cθ = (2-21b)
Difference in deflection s between two slab

Three types of foundation are included in this model, namely the Winkler foundation, the

semi-infinite elastic solid foundation and the layered elastic solid foundation. Three

contact conditions between slab and foundation can be considered: full contact, partial

contact without initial gaps, and partial contact with initial gaps.

(2) ILLI-SLAB

The two-dimensional finite element program ILLI-SLAB was originally developed at

the University of Illinois in 1977 for structural analysis of one- or two-layer concrete

pavements, with or without mechanical load transfer systems at joints and cracks

(Tabatabaie 1978). The original ILLI-SLAB model is based on the theory of a medium-

thick plate on a Winkler (dense liquid) foundation, and has the capability of evaluating

structural response of a concrete pavement system with joints and/or cracks. It employs

24
the 4-noded, 12-DOF plate bending element. The Winkler type subgrade is modeled as a

uniform, distributed subgrade through an equivalent mass formulation.

Since its development, ILLI-SLAB has been continually revised and expanded to

incorporate a number of options for support conditions, thermal gradient modeling

techniques, load transfer modeling techniques, material properties, and interaction

between the layers (contact modeling). Versions of this FEM program include ILLISLAB,

ILSL2, and the more recent interactive ISLAB2000.

Figure 2.5 shows the idealization of various components of the ILLI-SLAB model.

The rectangular plate element illustrated in Figure 2.6a is used to model the concrete slab

and base layer. There are three displacement components at each node: vertical

displacement (w) in the z-direction, rotation ( θ x ) about the x-axis and rotation ( θ y ) about

the y-axis (Tabatabaie, 1978).

In ILLI-SLAB, a dowel is simulated as bar element, as illustrated in Figure

2.6b.There are two displacement components at each node for a dowel bar: vertical

displacement (w) in the z-direction, and rotation ( θ y ) about the y-axis. A vertical spring

element is used to model the relative deformation of the dowel bar and the surrounding

concrete (Tabatabaie, 1978). Several subgrade models are available in the later versions of

ILLI-SLAB. In addition to the Winkler subgrade model, the program includes an elastic

solid foundation (Boussinesq model), two-parameter model (Vlasov), three-parameter

model (Kerr) and Zhemochkin-Siitsyn-Shtaerman formulations. Despite the options,

however, the Winkler foundation model is most often used due to its simplicity.

25
Figure 2.5 Finite Element Components Used in Development of Pavement System Model
in ILLI-SLAB (Tabatabaie and Barenberg 1980)

The assumptions used by ILLI-SLAB regarding the concrete slab, stabilized base,

overlay, dowel bars, keyway and aggregate interlock are briefly summarized as follows:

1. Small deformation theory of an elastic, homogenous medium-thick plate is employed

for the concrete slab, stabilized base and overlay. Such a plate is thick enough to carry

transverse load by flexure, rather than in-plane force (as would be the case for a thin

member), yet is not so thick that transverse shear deformation becomes important. In this

theory, it is assumed that lines normal to the middle plane in the undeformed state remain

straight, unstretched, and normal to the middle plane of the deformed plate. Each lamina

parallel to the middle surface is in a state of plane stress, and no axial or in-plane shear

stress develops due to loading.

2. In the case of a bonded stabilized base or overlay, full strain compatibility is assumed at

the interface. For the unbonded case, shear stresses at the interface are neglected.

3. Dowel bars at joints are linearly elastic, and are located at the neutral axis of the slab.

4. When aggregate interlock or a keyway is specified for load transfer, load is transferred

26
from one slab to an adjacent slab by shear. However, with dowel bars some moment as

well as shear may be transferred across the joints.

Various types of load transfer systems, such as dowel bars, aggregate interlock or a

combination of these can be considered at the slab joints and cracks. The model can also

accommodate the effect of another layer such as a stabilized base or an overlay, either

with perfect bonding or no bond. Thus ILLI-SLAB provides several options that can be

used in analyzing the following design and rehabilitation problems (Ioannides et al. 1985):

1. Multiple wheel and axle loads in any configuration, located anywhere on the slab;

2. A combination of slab arrangements such as multiple traffic lanes, traffic lanes and

shoulders, or a series of transverse cracks such as in continuously reinforced concrete

pavements;

3. Jointed concrete pavements with longitudinal and transverse cracks with various load

transfer systems;

4. Variable subgrade support, including complete loss of support over any specified

portion of the slab;

5. Concrete shoulders with or without tie bars;

6. Pavement slabs with a stabilized or lean concrete base, or asphalt or concrete overlay,

assuming either perfect bonding or no bond between the two layers;

7. Concrete slabs of varying thicknesses and moduli of elasticity, and subgrades with vary

moduli of support;

8. A linear or nonlinear temperature gradient in uniformly thick slabs; and,

9. Partial contact of the slab with the subgrade, with or without using an iterative scheme.

The ILLI-SLAB model has been extensively verified by comparison with the

available theoretical solutions and the results from experimental studies (Tabatabaie and

27
Barenberg 1980 and Ioannides 1984).

(3) WESLIQUID and WESLAYER

In the computer models WESLIQUID and WESLAYER (Chou and Huang

1979,1981), a concrete pavement system with joints is represented by two slabs, and the

slab is treated as a single rigid layer. As in KENSLAB and ILLI-SLAB, rectangular thin-

plate elements with three degrees of freedom per node are used to represent the slab. At

the joint, this model considers shear and moment transfer. Shear transfer is specified in

two options: a) efficiency of shear transfer which is defined as the ratio of vertical

deflections along the joint between the unloaded and loaded slabs, and b) diameter and

spacing of dowel bars. Two foundation models are considered, the Winkler foundation

and the semi-infinite elastic solid foundation. Two contact conditions between slab and

foundation are considered: full contact and partial contact.

(4) FEACONS

The FEACONS (Finite Element Analysis of CONcrete Slabs) program was

developed at the University of Florida and has been used in the analysis of existing

concrete pavements on a test road in Florida. In the FEACONS program, a jointed

concrete pavement is modeled by a three-slab system. A concrete slab is modeled as an

assemblage of rectangular plate bending elements with three degree of freedom at each

node. The three independent displacements at each node are (1) lateral deflection, (2)

rotation about the x-axis, and (3) rotation about the y-axis. The corresponding forces at

each node are (1) the downward force, (2) the moment in the x direction, and (3) the

moment in the y direction.

Load transfers across the joints between two adjoining slabs are modeled by shear

and torsional springs connecting the slabs at the nodes of the elements along the joint. The

28
subgrade is modeled as a Winkler foundation by a series of vertical springs at the nodes.

Subgrade voids are modeled as initial gaps between the slab and the springs at the

specified nodes. A spring stiffness of zero is used when a gap exists.

FEACONS III, the third version of FEACONS program, can analyze the response of

a concrete pavement system subjected to combinations of concentrated and uniform

applied vertical loads. The present version or fourth version of the program, named

FEACONS IV, was developed for analysis of plain jointed concrete pavements subjected

to load and temperature differential effects.

Three-dimensional models

Until the beginning of 1990s many 2-D assumptions were inevitably used in

pavement analysis due to limited computing power. As a result, many 2-D assumptions

tax the accuracy of numerical results predicted by FEM because they increase the

magnitude of approximation error by limiting natural behavior of a pavement system. On

the contrary, 3-D finite element approach requires less significant assumptions so that the

magnitude of approximation error can be reduced. In recent years, 3-D finite element

analysis (Davids et al. 1998; Brill 1998) has been applied to concrete pavement structure

analysis.

Figure 2.6 shows the concept for three different finite element analysis approaches

for the same pavement problem. Figure 2.5a shows a 3-D finite element approach using

tri-linear continuum solid finite elements with 8 nodes per element and 3 DOF per node.

Figure 2.5b demonstrates a 2-D axisymmetric analysis approach with 2-D bilinear

continuum solid elements, having 4 nodes per element and 2 DOF per node. Figure 2.5c

illustrates an approach using 2-D plates on an elastic foundation with Kirchhoff plate

bending elements, which have 4 nodes per element and 3 DOF per node. Although all

29
three approaches simulate rigid pavement systems, the formulations behind them are not

the same.

The axisymmetric 2-D approach has been one of the common ways to analyze flexible

pavement problems since Burmister created a layered linear elastic half-space solution in

1940s. It can also be used to analyze a multi-layered rigid airport pavement system. Under

the axisymmetric assumption, we can simulate multi-layered rigid pavement system with

distinct layer properties and apply nonlinear constitutive models. In the finite element

context, the largest benefit of this approach is the small problem size because we need to

model only a 2-D plane with axisymmetric elements that have 4 nodes and 2 DOF per

node. In addition, mesh construction is much easier than for a 3-D analysis because we

only need to create mesh for a 2-D plane. The first formulation of an axisymmetric model

for analysis of pavement structure was made by Wilson (1969, 1970). Despite the

advantages of these models for analysis of layered systems, they are not capable of

evaluating various concrete pavement features such as the finite dimensions of slab,

pavement joints, and various loading conditions such as edge or corner loading.

Continuum 3-D solid elements have been used to simulate both concrete slab and

supporting layers. As a result, there is no need of any 2-D assumptions limiting the natural

behavior of pavement systems. The behavior of concrete slab can be accurately simulated

with the 3-D finite element approach because the kinematic hypothesis is no longer

imposed and shear deformation is allowed for the continuum solid element.

30
Figure 2.6 Comparison Between 3-D And 2-D FEM Pavement Analysis Approach

Hence, nonlinear and shear deformations, created by the heavy multiple wheel loads and

temperature differential acting on a thick concrete slab, can be accurately simulated in

concrete slab modeling. The nonlinear and heterogeneous nature of rigid airport

pavements can be considered in the 3-D finite element approach by using appropriate

nonlinear constitutive models for each distinct layer of pavement systems. Detailed stress

and strain distributions can be investigated over the whole pavement system because

31
continuum solid elements are used in 3-D finite element analysis approach. Influence of

multiple wheel loads can be investigated in detail because non-axisymmetrically shaped

multiple pressure loads can be applied anywhere within the problem domain. However, 3-

D analysis may require a long computational time and large storage space due to the

problem size. The size of the 3-D problem is much larger than that of the 2-D one, because

of the extra DOF and large volume of the 3-D problem domain. For instance, consider a

uniform finite element mesh with elements along each axis. The size of the 2-D problem is

proportional to m 2 in plane, while that of 3-D problem is a function of m 3 in space. If the

number of elements is large, then the problem size becomes huge and quickly exceeds the

limit of available computational resource in common workstations or PCs. Hence, the

trade-off of 3-D finite element analysis is the long computational time and large data

storage space caused by the problem size.

EverFE (Davids et al. 1998) is a Windows-based three-dimensional (3D) rigid

pavement analysis tool, developed at the University of Washington in an attempt to make

3D finite element (FE) pavement analysis more accessible to users in a broad range of

settings. EverFE allows for simple and practical investigations of various factors (dowel

locations, gaps around dowels, temperature effects, etc.) on the response of pavement

structures, and parametric studies to evaluate different design and retrofit strategies. The

program incorporates graphical pre- and post-processing capabilities tuned to the needs of

rigid pavement modeling and allows transparent finite element model generation,

innovative computational techniques for modeling joint transfer, and efficient multi-grid

solution strategies (Davids et al, 1998). These features permit realistic models with

complex geometry to be generated in a matter of minutes, and solutions to be obtained on

32
desktop personal computers in a reasonable amount of time.

EverFE allows the user to specify all the parameters of the problem interactively, with

immediate visual feedback. Its intuitive graphical user interface (GUI) allows for easy and

efficient entry of these parameters, and allows users to easily test different designs,

perform parametric studies, and analyze the as-built configuration.

EverFE permits the modeling of one or multiple slabs with transverse joints in any

orientation. Elastic base layers below the slab may be explicitly modeled, and the

foundation below the elastic base layers is treated as a dense liquid foundation. Extended

shoulder may also be modeled. Immediate visual feedback is provided to the user as

parameters and dimensions are changed.

In its current version, EverFE assumes that the slab and foundation are linearly

elastic. The foundation may be specified as “no tension”, a useful feature if no base layers

are considered and the effect of slab lift-off is of interest.

EverFE allows the user to quickly specify dowels placed in common patterns, such as

equally spaced along transverse joints or located only within the wheelpaths. Dowel bars

are represented in the model as embedded quadratic beam elements, a model developed by

Davids et al (1998). The dowel model is illustrated in Figure 2.7 and an example of the

dowel-slab interaction is shown in Figure 2.8.

All dowels are assumed to be located at mid-thickness of the slab and may be

specified as bonded or unbonded. In addition, dowel looseness may be modeled by

specifying a gap between the dowels and the slab. The gap is assumed to vary linearly

from maximum value at the face of the joint to zero at a specified distance along the

embedded portions of the dowel. Any other aspects of dowel location and embedment are

user-controlled with immediate visual feedback in the plan and elevation views of the

33
system.

Figure 2.7 Embedded Dowel Element (Davids et al. 1998)

Figure 2.8 Example of Dowel-Slab Interaction (Davids et al. 1998)

EverFE permits the modeling of linear aggregate interlock, which is simulated as a

16-noded, “zero thickness”, quadratic interface element meshed between two quadratic

hexahedral elements. The elements are characterized by a stiffness value analogous to the

k-value in the Winkler foundation assumption. While this approach is computationally

34
convenient, it does not allow for the complex mechanism of aggregate interlock shear

transfer to be accurately modeled. Furthermore, like the Winkler model, it is difficult to

rationally select an appropriate spring stiffness.

In EverFE, the model allows detailed constitutive relations for shear transfer along

the aggregate interface to be incorporated into the finite element model. Figure 2.9

illustrates the essence of the model.

Figure 2.9 Distribution of Aggregate and Stresses on Spherical Particle (Davids et al. 1998)

Stresses are related by assumptions that the contact areas are about to slip, and thus:

τ = μfσ (2-22)

in whichτis the shear stress of the cement paste, σis the normal stress of the cement

paste, and μ f is coefficient of friction between the paste and aggregate.

Modeling the loss of contact between a slab and an unbonded base layer is critical

when considering temperature-induced curling. EverFE permits the user to model slab lift-

off and joint contact using a nodal contact approach (Figure 2.10). The slab and the base

layer are meshed separately but in a way that the locations of the bottom nodes of the slab

coincide with the locations of the top nodes of the base. Stress and displacement

conditions at each coinciding pair of slab/base nodes are monitored during the solution,

35
and the nodes are appropriately constrained or released. If no base layers are modeled, a

no-tension Winkler foundation may be specified directly below the slab to model the loss

of contact.

Figure 2.10 Contact Modeling in EverFE (Davids et al. 1998)


EverFE is capable of automatically generating a mesh. It produces hexahedral

elements for the slab and base layers, surface elements for the subgrade, and beam and

interface elements for modeling joint shear transfer. The user specifies the level of mesh

refinement and has control over solution techniques such as the choice to optimize

memory usage or solution time. Figure 2.11 shows a typical mesh generated by EverFE.

Figure 2.11 Finite Element Idealization of Two-Slab System in EverFE (Davids et al. 1998)

36
2.1.3.3 The Finite Difference Method (FDM)

Although it is a general consensus that the FEM has overwhelming advantages over

the FDM when applied to the analysis of pavement structures, the latter may be more

suitable or convenient to use in some cases. Since solutions to this class of problems (i.e.,

slab-on-grade) require a wealth of computer memory, and the FDM is known to utilize a

smaller amount of memory than the FEM, it is likely that the FDM

technique may be particularly useful in problems requiring large computer effort

(Ioannides 1984).

The FDM in its application to the slabs-on-grade problem replaces the governing

differential Equation and the boundary conditions by finite difference Equations. These

Equations describe the variation of the primary variable (i.e., deflection) over a small but

finite spatial increment. The most important criterion that governs the adequacy of the

finite difference approximation is the level of refinement of the finite difference grid.

2.1.3.4 Numerical Integration Techniques

The fourth category of computerized numerical techniques includes solutions

involving integrals of Bessel, elliptical or other functions over infinite and finite ranges.

This approach is conceptually different from the methods discussed previously. In the

FEM and the FDM, the numerical procedure begins with the governing differential

Equations and is thus an essential part of the final solution. On the other hand, numerical

integration techniques are a choice of how to evaluate the integrals to derive an expression

after considerable manipulation of the governing differential Equations and the boundary

conditions (Ioannides 1984).

2.1.3.5 Fourier Series Method

The solution by Navier was the first solution for the bending problem of simply

37
supported rectangular plates using double trigonometric series. This is a significant

success of applying Fourier series method in mechanics problems. In the classical theory

of elastic bodies, Fourier series method has long been used to find solutions for various

problems where polynomial solutions of the differential Equations to express the

mechanics problem cannot be found because of the complex properties of the input

parameters. Before the matrix analysis method was developed in structural mechanics, the

Fourier series method was often used to find solutions for various structural mechanics

problems.

The Fourier series method is a mathematical method to find solutions of complex

differential Equations set (Henwood et al. 1981). In practice, after the Fourier series

solutions of a differential Equations set are found, the approximate answer of the problem

can be obtained by truncating the series to a sum of finite number of items. This is an

approximate calculation method, and is different with numerical analysis method, such as

the finite element analysis method.

2.2 Analysis of the Effect of Temperature Distribution on Responses of

Concrete Pavements

2.2.1 Analytical Solutions

2.2.1.1 Analytical Solutions Based on Linear Temperature Profile

By assuming a linear temperature distribution across the slab thickness, Westergaard

(1926a) developed solutions for three cases. In case one, the slab is assumed to be infinite

in both x and y directions. The maximum warping stress is given by:

Eα t Δ t
σ 0= (2-23)
2(1 − μ )

38
where α t is the coefficient of thermal expansion of concrete, Δ t is the temperature

difference between the top and bottom surface of the pavement slab.

In case two, the slab is assumed to be infinite in positive y and both positive and

negative x directions, the maximum warping stresses at the top surface of slab in x and y

directions are given by:

⎡ ⎛ y π ⎞ ⎜⎜⎝ L 2 ⎟⎟⎠ ⎤
⎛ −y ⎞

σ y = σ 0 ⎢1 − 2 sin ⎜ + ⎟e ⎥ (2-24a)
⎢⎣ ⎝L 2 4⎠ ⎥⎦

⎡ ⎛ x π ⎞ ⎜⎜⎝ L 2 ⎟⎟⎠ ⎤
⎛ −x ⎞

σ x = σ 0 1 − μ 2 sin⎜
⎢ + ⎟e ⎥ (2-24b)
⎢⎣ ⎝L 2 4⎠ ⎥⎦

In case three, the slab is assumed to be infinite in both positive and negative x

directions, and have a finite width b along y direction. The maximum warping stress at the

top surface of the slab in y direction is given by:

⎧ 2 cos λ cosh λ ⎡
σ y = σ 0 ⎨1 − ⎢(tan λ + tanh λ ) cos y cosh y
⎩ sin (2λ )sinh (2λ ) ⎣ L 2 L 2

y ⎤⎫
+ (tan λ − tanh λ )sin
y
sinh ⎥⎬ (2-25)
L 2 L 2 ⎦⎭

b
where λ = . The corresponding warping stresses in x direction is
L 8

σ x = σ 0 + μ (σ y − σ 0 ) (2-26)

Bradbury (1938) developed a solution for warping stresses in a concrete pavement

slab with finite length a and width b based on Westergaard’s results. The maximum

warping stresses at the center of the slab are given by:

Eα t Δ t
(σ x )center = (C x + μC y )
(
2 1− μ 2 ) (2-27a)

39
Eα t Δ t
(σ ) = (C y + μC x )
y center
(
2 1− μ 2 ) (2-27b)

and the warping stress at the middle point of the longitudinal slab edge is given by:

Eα t Δ t
(σ x )edge = Cx (2-27c)
2

where

2 cos λ a cosh λ a ⎡
Cx = 1 − ⎢ (tan λ a + tanh λ a ) cos x cosh x
sin (2λ a )sinh (2λ a ) ⎣ L 2 L 2

x ⎤⎫
+ (tan λ a − tanh λ a )sin
x
sinh ⎥⎬ (2-28a)
L 2 L 2 ⎦⎭

2 cos λb cosh λb ⎡
Cy = 1− ⎢ (tan λb + tanh λb ) cos y cosh y
sin (2λb )sinh (2λb ) ⎣ L 2 L 2

y ⎤⎫
+ (tan λb − tanh λb )sin
y
sinh ⎥⎬ (2-28b)
L 2 L 2 ⎦⎭

a b
where λ a = and λb = .
L 8 L 8

2.2.1.2 Analytical Solutions Based on Nonlinear Temperature Profile

In the past, researchers tended to neglect the effect of the nonlinear temperature

distribution. Richardson and Armaghani (1988) concluded from an analysis of recorded

temperature data that the nonlinear temperature distribution in concrete pavement did not

have significant impact on its performance. However, other researchers suggested that the

effect of nonlinear temperature distribution on maximum warping stresses in concrete

pavements should not be neglected. Choubane and Tia (1992, 1995) carried out field

investigation and analysis on the effects of nonlinear temperature distribution on concrete

pavement slabs. The research was done on a six-slab experimental concrete pavement

40
constructed in Florida. They concluded that compared with the results obtained by

analysis assuming a linear distribution assumption, analysis based on a nonlinear

temperature profile produced lower maximum tensile stresses for the daytime condition

and higher for the nighttime condition. Their calculations showed that by considering the

nonlinear temperature effect, the calculated maximum tensile stresses in concrete

pavement could be 11 percent larger during the period from late night to early morning

time. The maximum tensile stress computed from recorded temperature data without

considering the effect of nonlinear temperature distribution was 216 psi, while the

computed result was 240 psi when the effect of nonlinear distribution was considered.

Mohamed and Hansen (1997) developed an analytical model to study the effect of

nonlinear temperature distribution. By applying their model to analyze recorded field

temperature data from research in Illinois and Florida and comparing the results with

those obtained based on assumption of linear temperature variation, they concluded that

analysis based on linear temperature distribution was limited by the assumption that

stresses at top and bottom surfaces of the slab are equal in value but opposite in sign.

Their comparison indicated that during late evening and early morning hours, the analysis

based on linear temperature profile underestimated the tensile stress at the bottom of

pavement slabs by a factor of 3. They also found that there is tensile stress occurred at the

top surface of the pavement slab by analysis based on nonlinear temperature profile. This

was in contradiction to prediction by the traditional linear analysis that there should be

compressive stress at slab top during this period. They thus concluded that the real

temperature distribution along the slab depth was important in the analysis of the effect of

temperature variation.

Zhang (2001) developed a closed-form model to analyze the effect of nonlinear

41
temperature distribution on warping stresses in concrete pavements. This model took into

consideration the exact slab dimensions, the effect of transverse shear deformation, and

the effect of subgrade inter-locking action. Similar findings to those by Mohamed and

Hansen (1997) and Choubane and Tia (1992) were reported.

2.2.2 Numerical Solutions

Finite element models, such as KENSLABS (Huang 1985), WESLIQUID and

WESLAYER (Chou 1981), and FEACONS IV (Tia el at 1988) described earlier in Section.

2.1.3.2, can also be applied to analyze thermal warping stresses in concrete pavement.

Based on the basic structures of these models described previously in Section 2.1.3.2,

the function for analyzing thermal warping stresses are incorporated under the assumption

that temperature varies linearly. In analyzing thermal warping stresses, the models assume

that each slab acts independently and is not restrained by lubricated dowel bars. Huang

(1985) pointed out that this assumption is reasonable if all the adjoining slabs are of the

same size and thickness, and warp the same amount at corresponding points along the

joint. Similar to analysis for traffic induced stresses, partial contact between slab and

foundation is considered in these three models for thermal warping stresses analysis.

The program ILSL2, an extension of the widely-used finite element program ILLI-

SLAB, can be used to calculate deflections and stresses in jointed slab-on-grade

pavements, with or without load transfer systems. The program can also accommodate a

stabilized base or an overlay, by assuming either no bond or perfect bond between the two

constructed layers, which are both modeled as plates. Another distinguishing feature of

ILSL2 lies in its treatment of temperature curling effects. The program allows the

computation of deflections and stresses due to temperature variation through the thickness

of the slab. Either linear or nonlinear temperature distributions through the thickness of

42
the slab can be accommodated. Wheel loads and gaps underneath the slab can also be

accommodated when this option is used.

EverFE, the 3D FE software mentioned in Section 2.1.3.2, can also be used to

calculate temperature induced warping stresses as well as deflection caused by nonlinear

temperature distribution through the slab thickness. In the current version EverFE 2.22,

the temperature profile can be input as many as 4 points at the following locations: the top

surface, one third of slab depth, two third of slab depth, and the bottom surface

respectively, which permits more accurate analysis to be done.

2.3 Summary of Literature Review

For structural analysis of concrete pavement under traffic loading or temperature

gradient, both analytical and numerical models have been developed by researchers.

Most of the analytical models were based on thin-plate theory and analyzed single slab

response without considering the effect of joint load transfer. However, this type of models

cannot consider the effect of transverse shear deformation and effects of joints. To

overcome these shortcomings, some researchers developed concrete pavement models

based on thick-plate theory, but the effect of joint load transfer have not been thoroughly

addressed.

Numerical models based on finite element method have greatly facilitated the

analysis of structural responses of concrete pavements under the action of traffic loadings.

They can consider the real geometric features of pavement slabs, such as the finite slab

length and width, the effect of joint force transfer, and partial contact at slab and subgrade

interface. However, numerical models are not perfect not only because computer storage

43
requirements and computation time pose certain limitation on their use, but also because

they have also been developed based on the classical thin-plate theory. The effects of

transverse shear deformation on stresses and deflections of concrete pavements can be

significant, but these numerical models cannot take them into consideration.

To incorporate the effects of joint load transfer into the analysis of jointed concrete

pavement system, the researchers at the Center for Transportation Research of the

National University of Singapore (NUS) have developed a three-slab thick-plate model

(Shi 1995) and a five-slab thick-plate model (Zhang 2001) that are able to predict the

stresses, deflections and fatigue properties of concrete pavements under vertically applied

loads. The three-slab model can deal with joints in one direction only. The five-slab model

can consider the effects of transverse and longitudinal joints at the same time.

However, both of them did not consider the effect of adjacent corner slabs. Thus,

improvements can be made using a more realistic multiple-slab models that can

consider the interactions between the loaded slab and all adjacent unloaded slabs.

Temperature is another important factor influencing the functioning of concrete

pavements. Traditionally temperature effects were treated under the assumption that

temperature varies linearly through the thickness of a pavement slab. In recent years, the

effects of nonlinear temperature distribution have drawn the attentions of many

researchers. However, a common drawback that exists in the aforementioned analytical

solutions is that they all modelled concrete pavement as single slab in which load transfer

across joints was not considered. Finite element analyses by Chou (1981, 1984) have

shown that the joint shear transfer capability had a significant influence on the state of

stresses in the slab. It is thus proposed that a nine-slab and a six-slab thick-plate model be

developed in this research to analyze the effect of nonlinear temperature distribution on

44
warping stresses in concrete pavements. These two models will be developed for a system

of nine or six rectangular slabs resting on a Winkler foundation. The proposed models will

take into account of the effects of transverse shear deformation, the effects of nonlinear

temperature distribution across the slab thickness, and the effects of shear force transfer

across joints.

45

You might also like