You are on page 1of 40

GENERAL VISCOELASTIC SOLUTIONS FOR THE MULTILAYERED SYSTEMS

SUBJECTED TO STATIC AND MOVING LOADS

Since the linear elastic layer solution for the layered systems was developed in the 1940s, the

linear elastic layer analysis has been systemized and widely used for the design of roadway

pavements as a tool for evaluating their structural soundness. The primary assumption made in

the analysis is the layered system consisting of materials that are linear elastic, and hence, an

application of the elastic layer analysis to asphalt mixtures, which is a well-known viscoelastic

material, has been limited. Therefore, the intention of the study was to derive a viscoelastic

solution able to take into account the time- and rate-dependent nature of the viscoelastic

materials in the multilayered system. In this paper, a linear viscoelastic solution for the

multilayered system subjected to a cylindrical unit step (static) load was derived from the elastic

solution using the principle of elastic-viscoelastic correspondence and the numerical inversion of

Laplace transforms. The solution was then extended to simulating pavement responses subjected

to a moving load by employing the Boltzmann’s superposition principle. The soundness of

output from the viscoelastic solution was confirmed by comparing them to those of the Finite

Element Analysis (FEA). Compared to the time and effort required in FEA, the analysis based on

the viscoelastic solution was faster and the results more reliable. Therefore, it is expected that the

viscoelastic solutions derived in this study will be an effective tool for the design of flexible

pavements.

CE keywords: Layered systems, Elasticity, Viscoelasticity, Flexible pavements, Closed form

solutions

INTRODUCTION
Designing pavement structures that are safe, reliable, and economical requires both efficient use

of materials and assurance of structural soundness. Their combined effects on pavement

structures in terms of stress and/or strain have been evaluated by the use of linear elastic layer

analysis programs, such as Bisar (De Jong et al. 1973) and Kenlayer (Huang 1993). These were

developed based on the well-known linear elastic theory for layered systems (Burmister 1945).

With the advent of the modern electric digital computer, the elastic layer analysis has been

systemized and widely used for the analysis and design of flexible pavements.

For the past few decades, the stiffness of asphalt concrete mixtures used for roadway

design and construction has been commonly characterized by means of resilient modulus,

defined as the ratio of the applied stress to the recoverable strain (for example, Roque et al.

1997). The loading function for this test, typically consisting of a 0.1-second haversine curve

followed by a 0.9-second rest period, was aimed at simulating the pattern of traffic loading over

pavement structures. By treating the resilient modulus as an elastic modulus, the structural

soundness of pavement has been evaluated through the elastic layer analysis. The resilient

modulus, however, is not a fundamental material property, and hence, the concept of resilient

modulus has been subsequently diminished in the latest Mechanistic-Empirical Pavement Design

Guide (ARA., ERES Division 2004). Instead, the new design guide uses the complex modulus,

representing the creep behavior of asphalt mixtures, as the primary measure of asphalt mixtures

in the procedure for both new construction and rehabilitation projects. It is, however, noted that

the new design guide still uses the linear elastic layer analysis as the primary model for the

evaluation of pavement performance. It is obvious that if the model of materials is changed, then

the structural model that is able to take their behavior into account should be used.

2
Therefore, the study was undertaken to develop a structural analysis model that is able to

integrate the viscoelastic behavior of asphalt mixtures into the multilayered system. The paper

begins with an overview of the linear elastic layer solution. Then, the corresponding linear

viscoelastic solution for the multilayered system subjected to a cylindrical unit step load is

derived. Because the behavior of viscoelastic materials is rate and time dependent, meaning that

their response is significantly influenced by the speed of traffic loading, the viscoelastic solution

is extended to formulate the structural behavior of the multilayered pavement system subjected

to a moving load.

REVIEW OF ELASTIC SOLUTION FOR A MULTILAYERED SYSTEM

Figure 1 illustrates the geometry and loading conditions of a multilayered system. Burmister

(1945) derived the general solution for the two- and three-layered systems based on the three

primary assumptions for the materials—homogeneous, isotropic, and linearly elastic—used in

the layered systems. By employing the theory of elasticity for the three-dimensional problem in

axisymmetric coordinates (Love 1927), his derivation begins with an assumption of a stress

function, F, with the Bessel function, J, which satisfies the equations of equilibrium and

compatibility for the axisymmetric conditions as follows:

F = J 0 (mr )[ Ai (m) exp(mz ) - Bi (m) exp(- mz ) + C i (m) × z × exp(mz ) - Di (m) × z × exp(- mz )] (1),

to which the subscripts i were added for the purpose of extending the elastic solution to the

multilayered system. The coefficients in Equation (1) were considered as functions of m because

Burmister essentially derived the elastic solution for the surface load of -mJ0(mr).

The following stress and displacement equations are obtained by substituting the stress

function into the general stress and displacement equations for the asymmetric conditions. Note

3
that the superscript * is an indication of the stress or displacement due to the surface load, -

mJ0(mr), not due to the uniform, cylindrical load q distributed over a circular area of radius a.

The equations are of the following forms:

ì Ai (m) × m 2 exp(m × z ) + Bi (m) × m 2 exp(- m × z )ü


ï ï
s z* i ( z , r ) = -mJ 0 (mr )í- C i (m) × m(1 - 2m i - mz ) exp(m × z ) ý (2-1)
ï+ D (m) × m(1 - 2 m - mz ) exp(- m × z ) ï
î i i þ

ì Ai (m) × m 2 exp(m × z ) - Bi (m) × m 2 exp(- m × z ) ü


t rz* i ( z , r ) = mJ 1 (mr )í ý (2-2)
îC i (m) × m(2 m i + mz ) exp(m × z ) + Di (m) × m(2 m i - mz ) exp(- m × z )þ

ì Ai (m) × m 2 exp(m × z ) - Bi (m) × m 2 exp(- m × z )ü


(1 + m ) ï ï
w * i ( z, r ) = - i
J 0 (mr )í- C i (m) × m(2 - 4 m i - mz ) exp(m × z ) ý (2-3)
Ei ï- D (m) × m(2 - 4 m + mz ) exp(- m × z ) ï
î i i þ

* (1 + m i ) ì Ai (m) × m 2 exp(m × z ) + Bi (m) × m 2 exp(- m × z ) ü


u i ( z, r ) = - J 1 (mr )í ý (2-4)
Ei î+ C i (m) × m(1 + mz ) exp(m × z ) - Di (m) × m(1 - mz ) exp(- m × z )þ

where s z* i ( z , r ) is vertical stress, t rz* i ( z , r ) is shear stress, w *i ( z , r ) is vertical displacement, and

u *i ( z , r ) is horizontal displacement. Radial stress s r* i ( z , r ) , radial strain e r* i ( z , r ) , and vertical

strain e z* i ( z , r ) , which will be used in the subsequent sections, are also given as:

é ì Ai (m) × m 2 exp(m × z ) + Bi (m) × m 2 exp(-m × z )ü ù


ê ï ï ú
êmJ 0 (mr )í+ C i (m) × m(1 + 2m i + mz ) exp(m × z ) ý ú
* ê ï- D (m) × m(1 + 2m - mz ) exp(-m × z ) ï ú
s r i ( z, r ) = ê î i i þ ú
ê 2 2
J (mr ) ì Ai (m) × m exp(m × z ) + Bi (m) × m exp(-m × z ) üú
ê- m 1 í ýú
êë mr î+ C i (m) × m(1 + mz ) exp(m × z ) - Di (m) × m(1 - mz ) exp(-m × z )þúû

(2-5)

ì Ai (m) × m 2 exp(m × z ) + Bi (m) × m 2 exp(- m × z )ü


(1 + m i ) æ J 0 (mr ) - J 1 (mr ) öï ï
e r* i ( z , r ) = - mç ÷í+ C i (m) × m(1 + mz ) exp(m × z ) ý (2-6)
Ei è 2 øï ï
î- Di (m) × m(1 - mz ) exp(- m × z ) þ

4
ì Ai (m) × m + Bi (m) × m × exp(-2m × z ) ü
* (1 + m i ) 2 ï ï
e z i ( z, r ) = - J 0 (mr )m exp(mz )í- C i (m) × (1 - 4 m i - mz ) ý (2-7)
Ei ï+ D (m) × (1 - 4 m + mz ) × exp(-2m × z )ï
î i i þ

The equations above are the stress, strain, and displacement equations for the elastic multilayered

system in which Ai(m) and Ci(m) are equal to zero at the lowest layer of n to satisfy the boundary

conditions of the multilayered system, as described in Figure 1.

In order to determine the coefficients, Ai(m), Bi(m), Ci(m), and Di(m), the sets of

simultaneous equations satisfying all the boundary conditions of the multilayered system need to

be solved. For the efficiency of computation analysis, the simultaneous equations are commonly

expressed in matrix form. For the continuity conditions between the layers, the coefficients

satisfying the boundary conditions are derived as follows:

· At the surface, z =0, sz* = -mJ0(mr) and the shear stress trz* is zero. By substituting these

conditions into Equations (2-1) and (2-2), a set of the simultaneous equations satisfying those

boundary conditions can be expressed in matrix form as:

é A1 (m) ù
ê B ( m) ú 1
[M 1 (m)] × ê 1 ú = éê ùú (3-1)
ê C1 (m) ú ë0û
ê ú
ë D1 (m)û

where,

m m - (1 - 2 m1 ) (1 - 2 m1 )ù
[M 1 (m)] = éê (3-2)
ëm -m ( 2 m1 ) (2 m1 ) úû

· For the continuous boundary conditions at the interfaces at z = h1, h2…hn, the vertical stress,

shear stress, vertical displacement, and horizontal displacement of the upper layer must be

the same as those of the lower layer. A series of simultaneous equations satisfying those

5
boundary conditions at interfaces between the upper and lower layers can be also expressed

in the following matrix notation:

é Ai (m) ù é Ai +1 (m) ù
ê B ( m) ú ê B ( m) ú
[M i (m)] × ê i ú = [M i +1 (m)] × ê i +1 ú (4-1)
ê C i ( m) ú ê C i +1 (m) ú
ê ú ê ú
ë Di (m)û ë Di +1 (m)û

where,

[M i (m)] =
ém exp(m × hi ) m exp(- m × hi ) - (1 - 2 m i - mhi ) exp(m × hi ) (1 - 2 m i - mhi ) exp(- m × hi ) ù
êm exp(m × h ) - m exp( - m × h ) ( 2 m + mh ) exp( m × h ) ( 2 m - mh ) exp( - m × h ) ú
ê i i i i i i i i ú
ê (1 + m i ) (1 + m i ) (1 + m i ) (1 + m i ) ú
ê- m exp(m × hi ) m exp(- m × hi ) (2 - 4 m i - mhi ) exp(m × hi ) (2 - 4 m i + mhi ) exp(- m × hi )ú
ê Ei Ei Ei Ei ú
ê (1 + m i ) (1 + m i ) (1 + m i ) (1 + m i ) ú
ê- m exp(m × hi ) - m exp(- m × hi ) - (1 + mz ) exp(m × hi ) (1 - mz ) exp(- m × hi ) ú
ë Ei Ei Ei Ei û

(4-2)

· Because the stresses and displacements must disappear as z approaches infinity, ¥, the above

matrix at the lowest layer n becomes:

é An -1 (m) ù
ê B ( m) ú B ( m)
[M n-1 (m)] × ê n-1 ú = [M n (m)] × éê n ùú (5)
ê C n -1 (m) ú ë Dn ( m) û
ê ú
ë Dn -1 (m)û

· By successive multiplications, the coefficients for the lowest layer can be related to Equation

(3-1):

é Bn ( m) ù -1 é1 ù
[
ê D (m)ú = [M 1 (m)] × [M 2 (m)] × [M 3 (m)] × [M 4 (m)] L[M n (m)] × ê0ú
-1 -1
] (6)
ë n û ë û

6
· Once Bn(m) and Dn(m) at the lowest layer are obtained from the above, they are substituted

into Equation (5). Then, the coefficients of the upper layers are successively determined by

solving the series of matrices as follows:

é An -1 (m) ù
ê B ( m) ú
ê n -1 ú = [M (m)]-1 × [M (m)] × é Bn (m) ù (7-1)
ê C n -1 (m) ú n -1 n ê D ( m) ú
ë n û
ê ú
ë Dn -1 (m)û

é A1 (m) ù
ê B ( m) ú
ê 1 ú = [M (m)]-1 L[M (m)] × [M (m)]-1 × [M (m)] × é Bn (m) ù (7-2)
ê C1 (m) ú 2 n-2 n -1 n ê D ( m) ú
ë n û
ê ú
ë D1 (m)û

· To find the stresses, strains, and displacements caused by the cylindrical load of q distributed

over the circular area of radius a, the Hankel transform method is often employed (Huang

1993). If R* is the stress, strain, or displacement in Equation (2), the stress, strain, or

displacement R corresponding to the loading condition can be obtained through the Hankel

transform shown below:

¥
R( z, r ) *
R( z, r ) = q × a × ò J 1 (m × a )dm (8)
0
m

Derivation of a Viscoelastic Solution for the Multilayered System

In order to derive a viscoelastic solution for the multilayered system shown in Figure 2, the

elastic-viscoelastic correspondence principle was employed for materials that are homogeneous,

isotropic, and linearly viscoelastic. The elastic-viscoelastic correspondence principle states that

the form of a relationship between stress and strain for a linear elastic material is the same as the

Laplace transform of the relationship for the corresponding linear viscoelastic material. The

7
procedure of applying the elastic-viscoelastic correspondence principle to known elastic

solutions is well documented in Findley et al. (1976). The steps for applying the elastic-

viscoelastic correspondence principle to the elastic solution derived in the previous section are

described below:

· Take the Laplace transforms of the elastic stress, strain, and displacement equations shown

above in which the coefficient Ai(m), Bi(m), Ci(m), and Di(m) are dependent on time because

they include the time-dependent material properties E(t) and m(t), and replace E and m with

sEˆ ( s ) and smˆ ( s ) where a caret (Ù) over the symbols means that the quantity has been

transformed and is a function of the Laplace variables, s. The Laplace transforms of the

equations are:

ì Aˆ i (m, s ) × m 2 exp(m × z ) + Bˆ i (m, s ) × m 2 exp(- m × z )ü


ïï ïï
sˆ z* ( z , r , s ) = - mJ 0 (mr )í- Cˆ i (m, s ) × m(1 - 2 smˆ i ( s ) - mz ) exp(m × z ) ý (9-1)
i
ï ˆ ï
+ Di (m, s ) × m(1 - 2 smˆ i ( s ) - mz ) exp(- m × z )
îï þï

ì Aˆ i (m, s ) × m 2 exp(m × z ) - Bˆ i (m, s ) × m 2 exp(- m × z )ü


ïï ïï
tˆrz* ( z , r , s ) = mJ 1 (mr )í+ Cˆ i (m, s ) × m(2 smˆ i ( s ) + mz ) exp(m × z ) ý (9-2)
i
ï ˆ ï
ïî+ Di (m, s ) × m(2 smˆ i ( s ) - mz ) exp(- m × z ) ïþ

ì Aˆ i (m, s ) × m 2 exp(m × z ) - Bˆ i (m, s ) × m 2 exp(-m × z )ü


(1 + smˆ i ( s )) ïï ïï
wˆ * i ( z , r , s ) = - J 0 (mr )í- Cˆ i (m, s ) × m(2 - 4smˆ i ( s ) - mz ) exp(m × z ) ý (9-3)
sEˆ i ( s ) ï ˆ ï
ïî- Di (m, s ) × m(2 - 4smˆ i ( s ) + mz ) exp(-m × z ) ïþ

ì Aˆ i (m, s ) × m 2 exp(m × z ) + Bˆ i (m, s ) × m 2 exp(- m × z )ü


(1 + smˆ i ( s )) ïï ïï
uˆ * i ( z , r , s ) = J 1 (mr )í+ Cˆ i (m, s ) × m(1 + mz ) exp(m × z ) ý (9-4)
sEˆ i ( s ) ï ˆ ï
ïî- Di (m, s ) × m(1 - mz ) exp(- m × z ) ïþ

8
é ì Aˆ i (m, s ) × m 2 exp(m × z ) + Bˆ i (m, s ) × m 2 exp(- m × z )ü ù
ê ïï ïï ú
êmJ 0 (mr )í+ Cˆ i (m, s ) × m(1 + 2 smˆ i ( s ) + mz ) exp(m × z ) ý ú
ê ï ˆ ï ú
ê ïî- Di (m, s ) × m(1 + 2 smˆ i ( s ) - mz ) exp(- m × z ) ïþ ú
sˆ r* ( z, r , s) = ê ú (9-5)
i
ê ì Aˆ i (m, s ) × m 2 exp(m × z ) + Bˆ i (m, s ) × m 2 exp(- m × z )üú
ê J (mr ) ïï ˆ ïïú
ê- m 1 í+ C i (m, s ) × m(1 + mz ) exp(m × z ) ýú
ê mr ï ïú
êë ïî- Dˆ i (m, s ) × m(1 - mz ) exp(- m × z ) ïþúû

ì Aˆ i (m, s ) × m 2 exp(m × z ) ü
ï ï
(1 + smˆ i ( s )) æ J 0 (mr ) - J 2 (mr ) öï+ Bˆ i (m, s ) × m exp(- m × z )
2
ï
eˆr* ( z, r , s) = - mç ÷í ý (9-7)
i
sEˆ i ( s ) è 2 øï+ Cˆ i (m, s ) × m(1 + mz ) exp(m × z ) ï
ï ˆ ï
î- Di (m, s ) × m(1 - mz ) exp(- m × z )þ

ì Aˆ i (m, s ) × m + Bˆ i (m, s ) × m × exp(-2m × z ) ü


(1 + smˆ i ( s )) ï
ï ïï
eˆ z* ( z, r , s) = - J 0 (mr )m 2 exp(mz )í- Cˆ i (m, s ) × (1 - 4 smˆ i ( s ) - mz ) ý
i
sEˆ i ( s ) ï ˆ ï
ïî+ Di (m, s ) × (1 - 4 smˆ i ( s ) + mz ) × exp(-2m × z )ïþ

(9-6)

· The same scheme outlined above should be applied to the simultaneous equations expressed

in the matrix form and also to the time-dependent material properties inside the matrices such

that:

é Aˆ1 (m, s ) ù
êˆ ú
B1 (m, s ) ú é1ù
[
ˆ ]
M 1 (m, s ) × ê =
ê Cˆ (m, s ) ú êë0úû
(10-1)
1
ê ú
êë Dˆ 1 (m, s )úû

where,

m - (1 - 2 smˆ 1 ( s )) (1 - 2 smˆ 1 ( s ))ù


[Mˆ (m, s)] = éêmm
1
-m (2 smˆ 1 ( s )) (2 smˆ 1 ( s ))úû
(10-2)
ë

and

9
é Aˆ i (m, s ) ù é Aˆ i +1 (m, s ) ù
êˆ ú êˆ ú
Bi (m, s ) ú Bi +1 (m, s ) ú
[ˆ ]
M i (m, s ) × ê
ê Cˆ (m, s ) ú
[
= Mˆ i +1 (m, s ) ]
× ê
ê Cˆ (m, s ) ú
(11)
ê i ú ê i +1 ú
êë Dˆ i (m, s )úû êë Dˆ i +1 (m, s )úû

where Aˆ n (m, s ) = Cˆ n (m, s ) = 0 at the lowest layer n.

· As shown in the elastic case, the coefficients for the lowest layer can be related to Equation

(10-1) as:

é Bˆ n (m, s ) ù é ù × é1ù (12)


-1

êˆ =[ ˆ
M
ú ê 1 ( m, s ) ][
× ˆ
M 2 ( m, s ) ] [
-1
× ˆ
M 3 ( m, s ) × ][
ˆ
M 4 ( m, s )
-1
L ˆ
M n] [
( m, s )
úû ê0ú
]
ë Dn (m, s )û ë ë û

· After Bn(m) and Dn(m) are obtained from the above equation, the coefficients of the layers

are solved in a successive manner:

é Aˆ n -1 (m, s ) ù
êˆ ú
ê Bn -1 (m, s ) ú = Mˆ (m, s ) -1 × Mˆ (m, s )
ê Cˆ (m, s ) ú
[ n -1 ] [n ]× éêDBˆˆ ((mm,, ss))ùú
n
(13-1)
ê n -1
ú ë n û
êë Dˆ n -1 (m, s )úû

é Aˆ1 (m, s ) ù
êˆ ú
ê B1 (m, s ) ú = Mˆ (m, s ) -1 L Mˆ (m, s ) × Mˆ (m, s ) -1 × Mˆ (m, s )
ê Cˆ (m, s ) ú
[2 ] [ n-2 n -1 ][ n ] [ ]× éêDBˆˆ ((mm,, ss))ùú
n
(13-2)
ê 1
ú ë n û
êë Dˆ 1 (m, s )úû

· All the coefficients of each layer are now functions of m and s. To find the stresses, strains,

and displacements caused by a cylindrical unit step load of q(t) = qH(t) distributed over the

circular area of radius a, where H(t) is the Heaviside step function, the Hankel transform is

applied to R*, and then the Laplace-transformed R̂ is obtained as:

10
Rˆ ( z , r , s ) *
¥
q
Rˆ ( z , r , s ) = × a × ò J 1 (m × a )dm (14)
s 0
m

Inversion of Laplace Transforms

Taking the inversion of the Laplace transforms is the final step in deriving a viscoelastic solution

for the multilayered system. One of the obstacles in deriving a viscoelastic solution by means of

the typical Laplace transform approach has been the complexity of inverting the Laplace

transforms, and hence, this study considered the numerical inversion technique to be used for

deriving the viscoelastic solution for the multilayered system. The numerical inversion of

Laplace transforms has been continuously improved with the advent of the modern electric

computer. There are nearly over 100 algorithms available for the numerical inversion. Among

them, this study used the inversion process of Laplace transforms, namely a Fixed Talbot

Algorithm (FTA), which was proposed by Abate and Valko (2004), because of its accuracy,

computational efficiency, and simplicity.


)
For the Laplace-transformed function f (s ) , the inverse transformation of f(t) is given by

the well-known inversion formula:

g + i×¥
1
f (t ) = × ò exp( s × t ) fˆ ( s )ds (15)
2p × i g-i ×¥

Many studies (for example, De Hoog et al. 1982) change Equation (15) to a form of Fourier

series consisting of the real and imaginary parts in order to make a better shape for numerical

analysis such that:

¥ ¥
exp(gt ) exp(gt )
f (t ) =
p ò [Re{fˆ ( p)}cos wt - Im{fˆ ( p)}sin wt ]dw = p ò [Re{fˆ ( p) exp(iwt )}]dw
0 0
(16)

where p = g+iw. The appearance of the above inversion formula looks familiar; however, it is

still not an effective function for numerical integration because of the infinite upper boundary. In

11
order to facilitate the numerical integration process, Abate and Valko (2004) introduced the

following form in which they converted the inversion formula to the complex domain based on

Talbot (1979) and deployed it by the Cauchy’s theorem:

f (t ) =
p
{ }
a (t ) éRe exp(t × a (t ) × q × (cot q + i )) × fˆ (a (t ) × q × (cot q + i )) ù
údq (17)
p ò0 ë´ {1 + i (q + (q cot q - 1) cot q )}
ê
û

2M
where a (t ) = and q is the angle in radians. As seen in the above, the upper limit is bounded
5t

as p, indicating that numerical integration would converge onto the expected value at a much

faster rate than that of the Fourier series approach (Equation (16)). The integral in Equation (17)

can be approximated by using the trapezoidal rule as:

é éì kp kp ü ùù
ê êïïexp(t × a (t ) × M × (cot M + i ))ïï úú
ê êí ý úú
a (t ) ê exp(a (t ) × t ) ˆ M -1
êï´ fˆ (a (t ) × kp × (cot kp + i )) ï úú
f (t ) = ê f (a (t )) + å Re êï ï úú (18)
M ê 2 î M M þ
k =1
ê úú
ê ê´ ì1 + i ( kp + æ kp cot kp - 1) cot kp öüú ú
ê ê í ç ÷ý
ë ë î M èM M M øþúû úû

where M, the number of terms to be summed, means the number of precision decimal digits and

is to control the round-off error in the computation of Equation (18). In other words, the higher

the number used, the more precision can be achieved, but the longer computation time is

required; therefore, the selection of a suitable number M is important in order to accomplish the

efficient computation analysis. The estimation of the computation error regarding the value of M

is given by Abate and Valko (2004) as:

f (t ) - f (t , M )
» 10 -0.6 M (19)
f (t )

The equation above indicates that the selection of M in the range between 5 and 10 would

provide a reasonable estimation of stress, strain, or displacement for most engineering problems.

12
Last, R( z , r , t ) is obtained by substituting Rˆ ( z , r , s ) in fˆ (s ) shown above. A complete form of the

inverse Laplace transformation of Rˆ ( z , r , s ) can be written as:

é é ì kp kp ü ùù
ê ê ïïexp(t × a (t ) × M × (cot M + i )) ïï ú ú
ê êReí ý úú
a (t ) ê exp(a (t ) × t ) ˆ M -1
ê ï´ Rˆ ( z , r , a (t ) × kp × (cot kp + i ))ï ú ú (20)
R( z, r , t ) = ê R ( z , r , a (t )) + å ê ï ïþ ú ú
M ê 2 î M M
k =1
ê úú
ê ê´ ì1 + i ( kp + æ kp cot kp - 1) cot kp öüú ú
ê ê í ç ÷ý
ë ë î M èM M M øþúû úû

By replacing R ( z , r , t ) with stress, strain, or displacement, any of these can be calculated at any

location and at any time.

VERIFICATION OF VISCOELASTIC SOLUTION

Pavement Geometry

For the axisymmetric condition, which was a primary basis of derivation of the viscoelastic

solution for the multilayered system, a pavement system consisting of four layers including the

rigid layer at its bottom and subjected to the circular unit step load with a uniform stress of 690

kPa and a contact radius of 152.4 mm was considered to be analyzed, thicknesses were assumed

to be typical values used for the design of pavements, and all layers were assumed to be fully

bonded, which is identical to the assumption made in the derivation above (Figure 3a). To

confirm the suitability of output from the viscoelastic solution derived in the previous section,

Finite Element Analysis (FEA) was chosen and performed for the same pavement structure but

with a finite width of 2.54 m. The width was estimated based on the previous work done by

Thompson (1982). The study recommended that the minimum distance of the width should be at

least 12 times the radius of the applied load. By introducing the fixed boundary conditions

13
underneath the subgrade layer (that is, this would act as the rigid layer), the three-layered

pavement system was prepared for FEA (Figure 3b).

From the literature, several studies (Al-Qadi et al. 2005, Elseifi et al. 2006, Kim et al.

2007, and Kim et al. 2009) used viscoelastic analysis, which considered the asphalt layer as

viscoelastic and the other layers as elastic using the Finite Element Method (FEM) based on a

two-dimensional or three-dimensional model for predicting the time-dependent responses of

flexible pavements. For the pavement structures (Figure 3), the same method was employed both

in the FEA and the analysis based on the viscoelastic solution. Except for the asphalt layer, the

linear elastic material properties for the base, subgrade, and rigid layers of the given pavement

structure were assumed to be the typical values used for the design of pavements (Table 1).

Material Input for Viscoelastic Analysis

This section describes the viscoelastic material properties used for the asphalt layer. Data

obtained from the previous studies was used as material input in both viscoelastic analyses. A

full description regarding laboratory testing can be found in the references cited below.

Lee and Kim (2009) and Kim and West (2010) proposed a methodology for determining

the viscoelastic creep compliance D(t) and Poisson’s ratio m(t) from the complex modulus test

using the indirect tension testing (IDT) mode. For the complex modulus test, field cores with a

diameter of 150 mm were cut to a thickness of 38 mm. Complex modulus tests were conducted

at various frequencies—0.01 Hz, 0.1 Hz, 0.5 Hz, 1 Hz, 5 Hz, 10 Hz, and 25 Hz—at a

temperature of 10ºC. In the studies, the authors concluded that the creep compliance and

Poisson’s ratio of asphalt mixtures can be represented by the well-known power functions as:

D(t) = D0+D1tm (21-1)

m1(t) = m0+m1tn (21-2)

14
where D0, D1, m, m0, m1, and n are regression coefficients. Taking the Laplace transformations

of the above equations yields:

D0 D1 × G(1 + m)
D( s) = + (22-1)
s s 1+ m

m0 m1 × G(1 + n)
m (s) = + (22-2)
s s 1+ n

Because creep compliance D(t) and relaxation modulus E(t) are two aspects of the same

viscoelastic behavior of materials, they are interchangeable using the following relationship

(Findley et al. 1976 and Tschoegl 1989):

1 1 s m -1
Eˆ ( s ) = 2 ˆ
= = m (23)
s D( s) D G(1 + m) s D0 + D1G(1 + m)
sD0 + 1 m -1
s

Equations (22-2) and (23) were used as input of the viscoelastic solution in Equations (9) and

(10), mˆ1 ( s ) and Eˆ 1 ( s ) . The values of the power model parameters, D0, D1, m, m0, m1, and n,

used are also presented in Table 1.

When performing FEA on a viscoelastic structure, most FE programs require the

viscoelastic shear and bulk moduli of given viscoelastic materials in forms of Prony series as

their input (Kim et al. 2010). The Prony series representation of creep compliance is of the

following form (Roque et al. 1997 and Kim et al. 2008):

N
ì t ü
D(t ) = D0 + å Di í1 - exp(- )ý (25)
i =1 î ti þ

where D0, Di, and h are Prony series parameters, and ti are retardation times. Since the Prony

series also can be used as an analytical representation of the time-dependent Poisson’s ratio of

viscoelastic materials (Kim and West 2010 and Kim et al. 2010), the time-dependent Poisson’s

15
ratio is, therefore, expressed as the identical form used in the creep compliance but with different

coefficients, m i:

N
ì t ü
m (t ) = m 0 + å m i í1 - exp(- )ý (26)
i =1 î ti þ

Lee and Kim (2009) and Kim and West (2010) also proposed a fitting technique for determining

the above Prony series from the power functions (Equation (21)). In fitting the creep compliance

and Poisson’s ratio using the Prony series shown in Equations (25) and (26), Prony series with

one spring element and six Kelvin elements (N=6) and retardation times with one decade interval,

ti = 10(i-4) (i = 1,¼ 6) were used, and then the unknown coefficients were determined by solving

the linear system of equations. According to the theory of linear viscoelasticity, the relationships

among the relaxation, D(t), shear, G(t), and bulk, K(t), moduli are expressed as algebraic

equations in the Laplace transform:

Ù 1
G (s) = Ù
(27-1)
2 s (1 + s m ( s )) Dˆ ( s )
2

Ù 1
K (s) = Ù
(27-2)
3s (1 - 2 s m ( s )) Dˆ ( s )
2

After substituting Laplace-transformed Equations (25) and (26) into Equation (27), taking the

inverse Laplace transforms of the above equations yields:

2 N -2
G (t ) = G0 + å G (exp(-r t ))
i =1
i i (28-1)

2 N -2
K (t ) = K 0 + å K (exp(-r t ))
i =1
i i (28-2)

The shear and bulk moduli determined through Equation (28) were used in FEA. The

coefficients of the shear and bulk moduli are present in Table 2.

16
Discretization of the FE model

In this study, the commercial FE software, ADINA 2008, was exclusively used. For the

axisymmetric stress condition, the rectangular element with nine nodes that has been

successfully used in two-dimensional FE analyses was used as the element type of model. A

typical meshing technique that achieves a finer mesh in the targeting area while keeping a

relatively coarser mesh in the remaining area was used as shown in Figure 3b. The numbers of

elements used for the asphalt, base, and subgrade layers were 696, 1044, and 2088, respectively.

The development of the time-dependent viscoelastic response depends not only on the

current state of stress or strain but also on the full history of its development. Considering the

nature of the numerical method (FE method), this indicates that time intervals between

computation points should be small enough to make an accurate prediction of the strain or stress

response. In order for the FE model to achieve such a condition, a loading time of 10 seconds

with an equal time interval of 0.01 seconds was applied. As a result, a total of 100 computation

points was analyzed.

Analysis Results

Because the radial stress or strain at the bottom of the asphalt layer and the vertical stress or

strain on the top of the subgrade layer have been considered critical responses and been used for

evaluation of cracking and rutting performance of flexible pavement systems, the stresses and

strains at each computation point compared to those from the viscoelastic solution at the same

time are presented in Figures 4 and 5 in which positive numbers means tension. Although a

slight gap between the values was observed, as we expected, because of the different nature of

FEM and the finite boundary conditions used for the FE model, not only the values but also the

trends were very close to each other. These observations appear to be sufficient for verifying the

17
soundness of the viscoelastic solution derived for the multilayered system; therefore, these lead

to the conclusion that the viscoelastic solution is a closed-form solution of the viscoelastic

multilayered system subjected to a circular unit step load.

As mentioned earlier, the value of M in Equation (18) could be determined in the range

between 5 and 10. Based on the estimation, a minimum value of 5 was used in the viscoelastic

analysis performed in order to expedite the computation process. It is, however, of interest to

know to what degree the assumption of the M value affects the stress and strain predictions.

Further evaluation was performed by varying the M values. Two extreme levels of M—5 and

10—were selected, and the same analysis with the M value of 10 was performed. Figures 6 and 7

show the stresses and strains at the bottom of the asphalt layer and at the top of the subgrade

layer in which the correlation in each case was evaluated by means of R2. For all cases, the R2

values were 1.00000. No difference between the values was observed visually or statically,

indicating that the M value of 5 is the suitable number of the summation in Equation 20.

FORMULATION FOR A MOVING LOAD

The viscoelastic solution derived in the preceding was extended to simulation of a moving load.

To formulate the responses of pavements to the moving load, the study used the Boltzmann’s

superposition principle (Findley et al. 1976 and Tschoegl 1989). For uniaxial loading conditions

shown in Figure 8, the Boltzmann’s superposition principle states that the sum of the strain

output resulting from each stress input is the same as the strain output resulting from the

combined stress input. Therefore, the strain output at any time tn corresponding to any arbitrary

stress input si can be obtained numerically from the following equation:

n
e (t n ) = s i × D(t n - t 0 ) + å [× s i D(t n - t i ) - s i -1 D(t n - t i -1 )], for t n = t1 Kt N (29)
i =1

18
Note that the same argument applies when step changes or arbitrary changes in strain are applied

and the resulting changes in stress are determined.

By employing the Boltzmann’s supposition principle, the viscoelastic solution was

extended to the derivation of a viscoelastic solution for the multilayered system subjected to a

moving load. The concept is illustrated in Figure 9. First, let us consider the moving load,

beginning at time t0 and distance r0 apart from the point P shown in Figure 9a. Suppose we

measure the responses of the given structure at the point P. At time tn, one of the structural

responses, defined as R ( z , r0 , t ) in the previous section, because of the load of qH(t) becomes

R( z , r0 , t n - t 0 ) similar to the first term in Equation (29). Figure 9b shows that the load was

shifted to the second location r1. At time tn, the responses at point P resulting from the combined

load input—one at time t0 and others at time t1—are summed similar to the strain at time t1 in

Equation (29). Then, the response at the point P becomes:

R ( z , rn , t n ) = R ( z , r0 , t n - t 0 ) + R ( z , r1 , t n - t1 ) - R ( z , r0 , t n - t1 ) (30)

If the above superposition principle is successively applied to multiple loading positions, ri, at

multiple times ti, one finds the following numerical integration:

n
R ( z , rn , t n ) = R ( z , r0 , t n - t 0 ) + å R ( z , ri , t n - t i ) - R ( z , ri -1 , t n - t i ) (31)
i =1

Given that time is related to the distance and speed of the moving load, the time ti is expressed as

follows:

xi -1 - xi
ti = + t i -1 (32)
V

where V is the speed of the moving load, and xi is the distance of the moving load from the point

P¢ in the loading axis, x, shown in Figure 10 (i.e. an independent loading axis should be used if

19
the point P is not placed in the loading axis). Assuming that time t0 at the beginning is zero, the

above can be expressed in a form of numerical integration as:

i x j -1 - x j
ti = å (33)
j =1 V

Substituting the above into Equation (31), the time variables can be removed and are replaced by

the distance variables as follows:

é x0 - x n ù
ê R ( z , r0 , )+ ú
x0 - x n ê V ú
R ( z , rn , )=ê n ì üïú
V x - xn i x j -1 - x j x0 - x n i x j -1 - x j
êå ïí R ( z , ri , 0 -å ) - R ( z , ri -1 , -å ) ýú
ê i =1 ï V j =1 V V j =1 V ïþúû
ë î

(34)

Note that because the distance from point P¢ could be a negative or positive value, the absolute

value of the difference between loading positions needs to be evaluated in Equation (34).

Finally, any structural response, such as stress, strain, or displacement, expressed in terms

of R at any loading position loaded at any speed can be obtained through Equation (34).

Application to Pavement Structure

For the pavement structure used above, a moving load starting from -2.54 m (r0 = x0 = -2.54 m)

apart from point P (=P¢ ) was shifted every 50.8 mm (ri-ri-1 = xi-xi-1 = 50.8 mm) to the loading

direction. A total 100 shifts was made to simulate one pass of the moving load. Figure 11a shows

a graphical description of the technique used in this analysis. Figure 11b shows the change in

tensile (radial) stress monitored at the bottom of the asphalt layer in the loading axis over one

pass of the moving load traveling at 60 km/h. As seen in the plot, the stress distribution is similar

to what we observe from the elastic analysis, whereas the stress distribution is not symmetric

around the peak stress because of the viscoelastic effect. Table 3 shows that the change in the

20
radial stress and strain at the bottom and the change in the vertical stress and strain on the top of

the subgrade layer at different speeds varied from 10 to 100 km/h. The stresses and strains were

monitored in the loading axis when the moving load passed point P (ri=xi=0). For all cases, the

tensile stress at the bottom of the asphalt layer increases, whereas the others decease as the speed

of loading increases. For the bottom stress and strain, the observations agree well with the

general behavior of viscoelastic materials where strain decreases under a fixed level of loading as

the applied loading frequency increases. For the top stress and strain at the subgrade layer, the

results also agree well with the general trend that the higher the asphalt layer modulus, the less

vertical strain and stress occur because of the reduction in deflection caused by the increase in

modulus.

These observations lead to the conclusion that the viscoelastic solution derived for the

moving load has a capability of simulating the pavement responses subjected to the moving load.

The benefit of the viscoelastic solution is that it significantly saves time and effort in evaluating

the structural responses of a pavement in comparison to those typically required to perform FEA

for the same structure. Therefore, it would be a fast and reliable tool for the design of flexible

pavement structure.

SUMMARY AND CONCLUSIONS

A general viscoelastic solution for the multilayered system was derived from the known elastic

solution using the principle of the elastic-viscoelastic correspondence and the numerical

inversion process of Laplace transforms. In order to take the time- and rate-dependent nature of

viscoelastic materials into account, the solution was then extended to simulation of the pavement

responses subjected to a moving load. The findings of this study are summarized as follows:

21
· To confirm the suitability of output from the viscoelastic solution for the cylindrical unit step

(static) load, a pavement system consisting of four layers including the rigid layer was

chosen, and the Finite Element Analysis (FEA) was performed for the same pavement

structure but with finite boundary conditions. Although a slight gap between the values was

observed because of the different nature of the FEA and the analytical approaches, the result

was very close and sufficient for verifying the soundness of the viscoelastic solution derived

for the multilayered system; therefore, it could be concluded that the viscoelastic solution

derived is a closed-form solution of the viscoelastic multilayered system subjected to a

cylindrical unit step load.

· The viscoelastic solution derived for the unit step loading case was extended to simulation of

pavement responses under a moving load. By employing the Boltzmann’s supposition

principle, the viscoelastic solution for static loading could be extended to the viscoelastic

multilayered system subjected to the moving load. From the simulation performed on the

same structure as that used in the verification study, it was found that the results reflected

well the general behaviors of viscoelastic materials, as well as those generally observed from

the elastic layer analysis. These observations lead to the conclusion that the viscoelastic

solution derived for the moving load has a capability of simulating the pavement responses

over the pass of the loading.

· Time and effort in evaluating the structural responses of a pavement through the viscoelastic

solutions derived in this study would not be comparable to those typically required to

perform FEA for the same structure. It is therefore expected that the viscoelastic solutions for

the multilayered system derived in this study will be a fast and reliable tool for the design of

flexible pavements.

22
REFERENCES

Abate, J. and Valko, P. P. (2004). “Multi-precision Laplace transform inversion,” International

Journal for Numerical Methods in Engineering, 60(21), 979-993.

ADINA user’s manual-version 8.5. (2008). ADINA R & D, Inc., Watertown, MA.

Al-Qadi, I. L., Yoo, P. J., Elseifi, M. A., and Janajreh, I. (2005). “Effects of tire configurations

on pavement damage.” Journal of Association of Asphalt Paving Technologists, 74, 921-

962.

ARA., ERES Division (2004). Guide for Mechanistic-Empirical Design of New and

Rehabilitated Pavement Structures. NCHRP 1-37A, Transportation Research Board,

National Research Council.

Burmister, D. M. (1945). “The general theory of stresses and displacements in layered soil

systems.” Journal of Applied Physics, 16, 89–94, 126–127, 296–302.

De Hoog, F. R., Knight, J. H., and Stokes, A. M. (1982). “An improved method for numerical

inversion of Laplace transforms,” Society for Industrial and Applied Mathematics, 3(3),

357-366.

De Jong, D. L., Peatz, M. G. F., and Korswagen, A. R. (1973). Computer Program BISAR,

Layered Systems under Normal and Tangential Loads. External Report. Koninklijke/Shell-

Laboratorium, Amsterdam, Netherlands.

Elseifi, M. A., Al-Qadi, I. L., and Yoo, P. J. (2006). ‘‘Viscoelastic modeling and field validation

of flexible pavements.” Journal of Engineering Mechanics ASCE, 132(2), 172–178.

Findley, W. N., Lai, J. S., and Onaran, K., (1976). Creep and relaxation of nonlinear viscoelastic

materials, Dover Publications, Inc., NY.

Huang, Y. H. (1993). Pavement Analysis and Design. Prentice-Hall, NJ.

23
Kim, J. and West, R. C. (2010). “Application of the viscoelastic continuum damage model to the

indirect tension test at a single temperature.” Journal of Engineering Mechanics ASCE,

136 (4), 496-505.

Kim, J., Lee, H., and Kim, N., (2010). “Determination of Shear And Bulk Moduli of Viscoelastic

Solids from the Indirect Tension Creep Test.” Journal of Engineering Mechanics ASCE,

in press.

Kim, J., Roque, R., and Byron, T., (2009). “Viscoelastic Analysis of Flexible Pavements and Its

Reological Effects on Top-Down Cracking,” Journal of Materials in Civil Engineering

ASCE, 21(7), 324-332.

Kim, J., Sholar, G., and Kim, S. (2008). ‘‘Determination of accurate creep compliance and

relaxation modulus at a single temperature for viscoelastic solids.” Journal of Materials

in Civil Engineering ASCE, 20(2), 147-156.

Kim, J., T. Byron, Sholar, G. A., and Kim, S. (2007). “Comparison of a Three-Dimensional

Viscoelastic Pavement Model to Full-Scale Field Tests,” Proceedings of the 86th Annual

Transportation Research Board, Washington, D.C.

Lee, H. and Kim, J. (2009). “Determination of viscoelastic Poisson’s ratio and creep compliance

from the indirect tension test.” Journal of Materials in Civil Engineering ASCE, 21(8),

416-425.

Love, A. E. H. (1927), Treatise on the Mathematical Theory of Elasticity, Dover Publications,

Inc., NY.

Roque, R., Buttlar, W. G., Ruth, B. E., Tia, M., Dickison, S. W., and Reid, B. (1997). Evaluation

of SHRP indirect tension tester to mitigate cracking in asphalt concrete pavements and

overlays. Final Report, FDOT B-9885, University of Florida, Gainesville, FL.

24
Talbot A. (1979) “The accurate numerical inversion of Laplace transforms.” Journal of the

Institute of Mathematics and Its Applications, 23, 97-120.

Thompson, M. R. (1982) ILLI-PAVE, User’s manual, Transportation Facilities Group,

Department of Civil Engineering, University of Illinois, Urbana.

Tschoegl, N. W. (1989). The phenomenological theory of linear viscoelastic behavior: an

introduction, Springer-Verlag, NY.

25
Table 1. Material input for the viscoelastic analysis

Power Model Parameters Elastic Material Properties


D0 (1/MPa) 4.840E-05 E2 (MPa) 207
D1 (1/MPa) 3.765E-04 E3 (MPa) 83
m 0.338 E4 (MPa) 1.E+20
µ0 0.143 µ2 0.40
µ1 0.231 µ3 0.45
n 0.008 µ4 0.01

26
Table 2. Coefficients in Prony series
Index, i ri (1/sec) Gi (MPa) ri (1/sec) Ki (MPa)
0 1.429E+02 4.520E+02
1 1.429E+03 2.131E+03 1.429E+03 2.477E+03
2 1.485E+02 1.491E+03 1.485E+02 2.286E+03
3 1.757E+01 1.305E+03 1.757E+01 2.025E+03
4 2.080E+00 8.261E+02 2.080E+00 1.487E+03
5 2.015E-01 3.843E+02 2.015E-01 7.494E+02
6 2.615E-02 1.997E+02 2.615E-02 4.610E+02
7 1.020E+03 -6.908E+00 9.096E+02 -1.776E+02
8 1.008E+02 -6.298E-01 9.600E+01 -2.525E+01
9 1.016E+01 -1.097E+00 9.172E+00 -5.166E+01
10 1.016E+00 -4.076E-01 9.095E-01 -2.732E+01
11 1.018E-01 -2.773E-01 8.862E-02 -2.588E+01
12 1.026E-02 -1.450E-01 8.055E-03 -2.385E+01

27
Table 3. Maximum stresses and strains for one pass of loading
At the Bottom of Asphalt Layer At the Top of Subgrade Layer
Speed (km/h) Radial Stress (kPa) Radial Strain Vertical Stress (kPa) Vertical Strain
10 2176 2.46E-04 -56 -6.00E-04
20 2327 2.27E-04 -54 -5.72E-04
30 2411 2.17E-04 -53 -5.57E-04
40 2469 2.10E-04 -52 -5.46E-04
50 2513 2.06E-04 -51 -5.38E-04
60 2582 1.98E-04 -50 -5.24E-04
70 2576 1.99E-04 -50 -5.26E-04
80 2601 1.96E-04 -50 -5.21E-04
90 2622 1.94E-04 -50 -5.17E-04
100 2640 1.92E-04 -49 -5.14E-04

28
Axis of Symmetry
a
q
r ¥
E1, m1 h1

E2, m2 h2
¼

hn

En, mn z

Figure 1. Geometry and loading conditions of an elastic multilayered system

29
Axis of Symmetry
a
q(t)

E1(t), m1(t) h1

E2(t), m2(t) h2
¼

hn

En(t), mn(t) z

Figure 2. Geometry and loading conditions of a viscoelastic multilayered system

30
Axis of Symmetry a = 152.4 mm
q(t) = 690×H(t) kPa
r ¥

Asphalt Layer 101.6 mm

Base Layer 304.8 mm

Subgrade Layer 1168.4 mm

Rigid Layer z ¥

(a)

Axis of Symmetry

(b)

Figure 3. Pavement structures for a) the viscoelastic solution analysis and b) the FEA

31
Viscoelastic Solution FEA

2500

2000
Stress (kPa)

1500

1000

500

0
0.0 2.0 4.0 6.0 8.0 10.0 12.0
Tim e (sec)

(a)
Viscoelastic Solution FEA

6.00E-04

5.00E-04

4.00E-04
Strain

3.00E-04

2.00E-04

1.00E-04

0.00E+00
0.0 2.0 4.0 6.0 8.0 10.0 12.0
Tim e (sec)

(b)

Figure 4. (a) Stresses and (b) strains at the bottom of the asphalt layer

32
Viscoelastic Solution FEA

0.0 2.0 4.0 6.0 8.0 10.0 12.0


0
-10
-20
Stress (kPa)

-30
-40
-50
-60
-70
-80
-90
Tim e (sec)

(a)
Viscoelastic Solution FEA

0.0 2.0 4.0 6.0 8.0 10.0 12.0


0.00E+00
-1.00E-04
-2.00E-04
-3.00E-04
-4.00E-04
Strain

-5.00E-04
-6.00E-04
-7.00E-04
-8.00E-04
-9.00E-04
-1.00E-03
Tim e (sec)

(b)

Figure 5. (a) Stresses and (b) strains at the top of the subgrade layer

33
M=5 M=10

2500
R2=1.00000
2000
Stress (kPa)

1500

1000

500

0
0.0 2.0 4.0 6.0 8.0 10.0 12.0
Tim e (sec)

(a)
M=5 M=10

6.00E-04

5.00E-04

4.00E-04
Strain

3.00E-04

2.00E-04

1.00E-04
R2=1.00000
0.00E+00
0.0 2.0 4.0 6.0 8.0 10.0 12.0
Tim e (sec)

(b)

Figure 6. (a) Stresses and (b) strains with different M values at the bottom of the asphalt layer

34
M=5 M=10

0.0 2.0 4.0 6.0 8.0 10.0 12.0


0
-10
R2=1.00000
-20
Stress (kPa)

-30
-40
-50
-60
-70
-80
Tim e (sec)

(a)
M=5 M=10

0.0 2.0 4.0 6.0 8.0 10.0 12.0


0.00E+00
-1.00E-04
-2.00E-04 R2=1.00000
-3.00E-04
Strain

-4.00E-04
-5.00E-04
-6.00E-04
-7.00E-04
-8.00E-04
-9.00E-04
Tim e (sec)

(b)

Figure 7. (a) Stresses and (b) strains with different M values at the top of the subgrade layer

35
si

sn
¼

s3
s2
s1
s0
t0 t1 t2 t3 ¼ tn

Figure 8. Illustration of Boltzmann’s superposition principle

36
a
q(t)
r, x
z
r0, t0
P
= R ( z , r0 , tn - t0 )

Loading Direction

¥
(a)
a a
q(t)
r, x
-q(t) z
r0, t0 R ( z , r0 , tn - t0 )
P =
- R ( z , r0 , tn - t1 )
r1, t1
+ R ( z , r1 , tn - t1 )

Loading Direction

z z

¥ ¥
(b)

Figure 9. Extended Boltzmann’s superposition principle to a moving load

37
Top View
r (Radial Axis)
Center of Loading P
ri

Loading Path x (Loading Axis)


xi P¢

Figure 10. Schematic illustration of the loading axis and point P¢

38
Top View

Axis of Symmetry
50.8 mm

P (=P¢)

r, x
Loading Path

(a)

3000

2500
Direction of Loading
2000
Stress (kPa)

1500

1000

500

0
-2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5
-500

-1000
Distance from Loading (m )

(b)

Figure 11. (a) Illustration of load shift; (b) stress distribution over one pass of loading

39

You might also like