Sie sind auf Seite 1von 31

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/241497958

ON THE RELATIONSHIP BETWEEN LOAD AND DEFLECTION IN RAILROAD


TRACK STRUCTURE

Article · January 2008

CITATION READS

1 2,362

5 authors, including:

Sheng Lu Shane Farritor


University of Nebraska at Lincoln University of Nebraska at Lincoln
8 PUBLICATIONS   26 CITATIONS    126 PUBLICATIONS   2,152 CITATIONS   

SEE PROFILE SEE PROFILE

Mahmood Fateh Gary Carr


U.S. Department of Transportation United States Department of Transportation, Washington DC
65 PUBLICATIONS   483 CITATIONS    31 PUBLICATIONS   106 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Onboard technology to detect changes in bridge behavior View project

Track Support Inspection View project

All content following this page was uploaded by Gary Carr on 19 August 2015.

The user has requested enhancement of the downloaded file.


ON THE RELATIONSHIP BETWEEN LOAD AND DEFLECTION IN RAILROAD TRACK STRUCTURE

Sheng Lu, Richard Arnold, Shane Farritor* Mahmood Fateh, Gary Carr
*Corresponding Author
Federal Railroad Administration
Department of Mechanical Engineering
Office of Research and Development
University of Nebraska Lincoln
1200 New Jersey Avenue SE
N104 Scott Engineering Center
Washington, DC 20590
Lincoln, NE 68588-0656

ABSTRACT

Track Modulus, defined as ratio between the rail deflection and the vertical contact pressure between the rail base

and track foundation, is an important parameter in determining track quality and safety. The Winkler model is a

widely used mathematical expression that relates track modulus to rail deflection. The Winkler model represents

railroad track as an infinitely long beam (rail) on top of a uniform, linear, and elastic foundation. The contact

pressure between the rail base and track foundation increases linearly with vertical deflection. However, it is widely

accepted that actual track deflection is highly non-linear. Several other models have been used to better represent

the behavior of railroad track structure including a model that includes a shear layer and one that uses discrete

supports.

This paper presents a new model of track deflection where the elastic foundation beneath the rail has a cubic

polynomial relationship between applied pressure and vertical deflection. This new cubic model is compared to

other models of railroad track structure, including the Winkler, Pasternak, and Discrete Support models, as well as

with experimental data. It is shown that the cubic model is a better representation of real track structure.

INTRODUCTION

Background

The relationship between applied loads, track stresses, and track deformations are important factors to be considered

in proper track design and maintenance. A representative mathematical model that accurately describes this

relationship is desirable. Winkler proposed the use of an elastic beam theory to analyze rail stresses and calculation

of a fundamental parameter, called the track modulus, which represents the effects of all the track components under

the rail (1). Track Modulus (represented by u in this paper) is defined as the supporting force per unit length of rail

per unit rail deflection (2). Track Stiffness (represented by k in this paper) is simply the ratio of applied load to
Lu et al. 2

resulting vertical deflection. Track stiffness relates load to deflection while track modulus relates a distributed load

to deflection.

Railway track has several components that all contribute to track modulus including the rail, subgrade,

ballast, subballast, ties, and fasteners. The rail directly supports the train wheels and is supported on a tie pad and

held in place with fasteners to crossties. The crossties rest on a layer of rock ballast and subballast used to provide

drainage. The soil below the subballast is the subgrade.

The subgrade resilient modulus and subgrade thickness have the strongest influence on track modulus.

These parameters depend upon the physical state of the soil, the stress state of the soil, and the soil type (3, 2).

Track modulus increases with increasing subgrade resilient modulus, and decreases with increasing subgrade layer

thickness (2). Ballast layer thickness and fastener stiffness are the next most important factors (2, 4). Increasing the

thickness of the ballast layer and or increasing fastener stiffness will increase track modulus (5, 2). This effect is

caused by the load being spread over a larger area. The system presented in this paper measures the net effective

track modulus that includes all these factors.

Track modulus is important because it affects track performance and maintenance requirements. Both low

track modulus and large variations in track modulus are undesirable. Low track modulus has been shown to cause

differential settlement that then increases maintenance needs (6, 7). Large variations in track modulus, such as those

often found near bridges and crossings, have been shown to increase dynamic loading (8, 9). Increased dynamic

loading reduces the life of the track components resulting in shorter maintenance cycles (9). It has been shown that

reducing variations in track modulus at grade (i.e. road) crossings leads to better track performance and less track

maintenance (8). Ride quality, as indicated by vertical acceleration, is also strongly dependent on track modulus.

The economic constraints of both passenger and freight rail service are moving the industry to higher-speed

rail vehicles and the performance of high-speed trains are strongly dependent on track modulus. It has been shown

that at high speeds there will be an increase in track deflection caused by larger dynamic forces (10, 11). These

forces become significant as rail vehicles reach 50 km/hr (30 mph) (12) and rail deflections increase with higher

vehicle speeds up to a critical speed (11). It is suggested that track with a high and consistent modulus will allow for

higher train speeds and therefore increased performance and revenue (11).
Lu et al. 3

An improved mathematical understanding of the relationship between loads and deflections will lead to

better track design and increased safety.

Problem Definition: A Beam on an Elastic Foundation (BOEF) Model of Track Structure

The BOEF model describes a point load applied to an infinite Bernoulli beam on an infinite elastic foundation.

Figure 1 shows a free load and deflection diagram of the rail under a one-wheel load (Figure 1, top). Here, the rail is

considered as a continuously supported beam where x represents the distance along the beam and w(x) represents the

vertical beam deflection. The approximation that the rail is continuously supported improves as the cross-tie

spacing decreases and as the rail bending stiffness increases (i.e. modulus of elasticity and second moment of area).

The applied load, P, is assumed to be a point load and creates a distributed load on top of the rail, p(x),

0+
where P = ∫ − p( x)dx . The supporting structure supports the bottom of the rail with a reaction distributed force, q(x).
0

In real track the supporting structure consists of tie plates, fasteners, cross-ties, ballast, etc. In the Winkler model

this supporting structure is an infinite elastic medium.

The difference in the vertical distributed force applied to the beam (q(x) and p(x)) causes curvature in the

beam as given by the following differential equation:

d 4w
EI = q( x) − p( x) , or (1)
dx 4

d 4w
EI + p ( x ) = q ( x)
dx 4

The solution to the differential equation is dependent upon the boundary conditions of the beam as well as the

loading conditions. A free body diagram that shows sections of the beam is shown in Figure 2. Here it can be seen

that one half the applied load the boundary condition for a concentrated applied load, P, must be supported by the

foundation reaction distributed force, q(x), on each half of the infinite beam, or:

∞ P
∫0
p( x)dx =
2
(2a)

In addition, symmetry and the stiffness of the beam demand that the slope of the beam be zero at the point of

loading.
Lu et al. 4

dw
=0 (2b)
dx x =0

The above differential equation and boundary conditions can now be set up and solved in different ways to represent

various track behavior. Four such solutions are defined in Section 3.

FIELD MEASUREMENTS OF TRACK MODULUS

Figure 3 shows the experimental results of the track responses under various applied loads. Rail deflection was

measured at given locations using linear variable differential transformers (LVDTs) as a short, slow moving train of

known weight passed. The axles of the train weighed 150600 N (33850 lbf), 60230 N (13540 lbf), and 30650 N

(6890 lbf). The LVDTs were mounted to steel rods (about 1m (3ft)) driven into the subgrade to provide a stable

reference. The LVDTs then measured the vertical motion of the flange relative to the steel rod. The results from

four LVDTs are shown in Figure 3. Here the LVDTs were placed at 1m (3ft) increments along the track (x=1m, 2m,

3m, 4m).

These measurements, along with many others dating back to the Talbot Report (13) clearly indicate that the

vertical rail deflections are not linearly proportional to the wheel loads. It is also important to note that the “degree”

on non-linearity can change dramatically over very short distances along the track. Note the deflection of the track

under the 30650 N (6890 lbf) load doubled over a distance of one meter. This non-linearity and variability greatly

complicates determining and modeling track structure. Several methods have been developed for calculating

modulus with each method assuming a different definition of track modulus that approximate the non-linear

behavior of real track.

Consider the definitions of track modulus represented in Figure 4 and described in the following sections.

Beam On an Elastic Foundation (BOEF) Method

The most straightforward method to estimate track modulus at a given track location is to simply measure the

vertical deflection at the point (w(0)=wo) of an applied known load, P. This is a measurement of the track stiffness,

k, but this measurement can be related to track modulus, u, using the BOEF model and assuming that the

relationship between rail supporting load p(x) and deflection w(x) is linear and elastic (i.e. p(x)=uw(x) as in 2, 1).

These assumptions lead to the Winkler model as described in Section 3.1. The resulting track modulus is given by:
Lu et al. 5

1 4
1 ⎛ 1 ⎞3 ⎛ P ⎞3
u = ⎜ ⎟ ⎜⎜ ⎟⎟ (3)
4 ⎝ EI ⎠ ⎝ w0 ⎠

where:

u is the track modulus

E is the modulus of elasticity of the rail

I is the moment of inertia of the rail

P is the load applied to the track

w0 is the deflection of the rail at the loading point

This method only requires a single measurement and it has also been suggested to be the best method for

field measurement of track modulus (14). However, as shown in Figure 3, it is clear that this linear approximation

has large error for real track. Using a single applied load and a single measurement of deflection does not capture

the changes in the load deflection curve present in real track.

Deflection Basin Method

The Deflection Basin Method uses the vertical equilibrium of the loaded rail and several deflection measurements to

more directly estimate track modulus. In this approach, rail deflection caused by a point load(s) is measured at

several (ideally infinite) locations along the rail and the entire deflected “area” calculated. This method requires

several deflection measurements over the section of track that supports the load(s), which makes it more time

consuming (2). Using a force balance this deflected area, or deflection basin, can be shown to be proportional to the

integral of the rail deflection (2, 1):

∞ ∞
P = ∫ q (x ) dx = ∫ uδ (x ) dx = uAδ (4)
−∞ −∞

where:

P is the load on the track

q(x) is the vertical supporting force per unit length

u is the track modulus

δ(x) is the vertical rail deflection

Aδ is the deflection basin area


Lu et al. 6

(area between the original and deflected rail positions)

x is the longitudinal distance along the track

These measurements and calculations result in a numerical solution to the BOEF equation given in Equation

(1). This solution does include the non-linear behavior of the track, but the measurements are extremely time

consuming and only reveal the track modulus at a given point. As shown in Figure 3, these measurements could

change dramatically for a point just centimeters away.

Heavy-Light Load Method

Many have represented the load/defection curve as piecewise linear with a low stiffness at low loads and a much

higher stiffness at higher loads (15). This is seen in real track as slack in the rail and can be caused by many things

such as the ties not contacting the ballast. As the rail is loaded, a low stiffness is experienced until the tie contacts

the ballast resulting in a higher stiffness. This leads to a measurement of track stiffness using two loads, Figure 4,

that are ideally both in the “high stiffness” range (e.g. slack is removed) (16, 6, 17).

P2 − P1
k=
w2 − w1
(5)
where:

k is the track stiffness

P1 and P2 are the applied loads

w1 and w2 are the corresponding

deflections

Again, a linear assumption is used to then transform the stiffness measurements of the two loads to track

modulus (substitute k = P into Equation 3). The clear difficulty with this measurement is that the real
wo

load/deflection relationship is not piecewise linear and the resulting stiffness varies with the selection of the two

loads, P1 and P2.

Track Modulus at Characteristic Load

It is proposed in this paper that a good definition of track modulus is the variation in supporting distributed force

relative to the variation in deflection near the characteristic load for a given track. This characteristic load might be
Lu et al. 7

defined as the nominal axle load for a given freight line (e.g. 160kN or 286,000/8=36kips). This can be expressed

mathematically as the derivative of the pressure deflection curve evaluated at the characteristic load P*:

∂p
u* = (6)
∂w *
P

where:

u is the track modulus

p is the supporting force per unit length of rail

P* is the characteristic load corresponding

for a given rail line

To evaluate the derivate at the characteristic load, the load must again be transformed to a distributed load.

This can be done with the linear assumptions as described above or with the cubic model given in Section 3.4. This

definition of track modulus has been used in field measurements (18).

MODELS OF RAIL DEFLECTION

The Winkler Model

In the Winkler model, the BOEF model described above assumes the distributed supporting force of the track

foundation is linearly proportional to the vertical rail deflection (i.e. p(x)=uw(x)). The BOEF differential equation

then becomes:

d 4 w( x)
EI + uw( x) = q( x) (7)
dx 4

This model has been shown to be an effective method for determining track modulus (19, 20) and

derivations can be found in (12, 21). The vertical deflection of the rail, w, as a function of longitudinal distance

along the rail x (referenced from the applied load) is given by:

Pβ − β ⋅ x
w( x ) = − e [cos(β x ) + sin (βx )] (8)
2u

where:
Lu et al. 8

1
⎛ u ⎞4
β =⎜ ⎟
⎜ 4 EI ⎟
⎝ ⎠

where:

P is the load on the track

u is the track modulus

E is the modulus of elasticity of the rail

I is the moment of inertia of the rail

w is the longitudinal distance along the rail

When multiple loads are present, the rail deflections caused by each of the loads are superposed (assuming

small vertical deflections) (21).

A plot of the rail deflection given by the Winkler model over the length of a four-axle coal hopper is shown

in Figure 5. The deflection is shown relative to the wheel/rail contact point for five different reasonable values of

track modulus (6.89, 13.8, 20.7, 27.6, and 34.5 MPa or 1000, 2000, 3000, 4000, and 5000 lbf/in/in). The model

assumes 115 lb rail with an elastic modulus of 206.8 GPa (30,000,000 psi) and an area moment of inertia of 2704

cm4 (64.97 in4).

The limitations of the Winkler model are clear given the widely accepted non-linearity of track structure.

However, this model is often used because it does provide a clear closed form solution to the relationship between

load and deflection in track structure.

Discrete Support Model

A second model assumes a similar linear relationship between the rail support and deflection, but uses discrete

springs to provide the rail support forces rather than the infinite elastic medium used in the Winkler model. The

discrete support model is similar to the Winkler model when the ties are uniformly spaced, have uniform stiffness,

and the rail is long. The discrete springs represent support at the crossties and the single applied load represents one

railcar wheel and is fully described in Norman (22).

The discrete support model is useful because track modulus can vary from tie to tie (as in Figure 3). The

proposed model also only considers finite lengths of rail and a finite number of ties, Figure 6. To reduce the model’s
Lu et al. 9

computational requirements, the rail is assumed to extend beyond the ties and is fixed at a (large) distance from the

last tie. This ensures the boundary conditions are well defined (the rail is flat, far away, or w(∞)=0 and w’(∞)=0)

and the rail shape is continuous, Figure 6 top.

The deflection in each of the springs (i.e. the rail deflection) can be determined by first solving for the

forces in each of the springs using energy methods and the free body diagrams in Figure 6. The principles of

stationary potential energy and Castigliano’s theorem on deflections are applied (21). These methods require small

displacements and linear elastic behavior. The number of equations needed to determine the forces in the springs is

equal to the number of springs (i.e. spring forces are the unknowns).

The moment and shear force in the cantilevered sections of the model (Figure 6(A) and (C)) can now be

calculated. Static equilibrium requires the moment, for Section A, to be:

M 1 = M A + V A x1 , and M 2 = M C + VC x2 (9)

Similar equations can be written for the sections of beam between each of the discrete supports. This leads to N+4

equations where N is the number of discrete supports used in the model. Now, the total system energy can be

written as the sum of the energy stored in the bending beam (sheer energy is negligible) and the energy stored in the

springs:

Mi F2
U TOTAL = U Beam + U Springs = ∑ ∫ dx + ∑ i (10)
2 EI 2 ki

Where Mi is the bending moment in each segment of the beam and E and I are the sectional properties. The

bending energy in each segment is summed. Also, Fi is the force in each support spring and ki is the stiffness of

each spring.

Castigliano’s theorem can now be used to create equations needed to solve for the unknown spring forces

and boundary conditions (moment and shear forces):

∂U ∂U ∂U ∂U ∂U
= = = 0, and = =0 (11)
∂Fi ∂M A ∂M B ∂VA ∂VB

With these relationships, a set of N+4 equations and N+4 unknowns can be developed by substituting the moment

expressions into Equation (12). These expressions can be written in matrix form:

(12)
Lu et al. 10

MF = P

where:

P is the load vector

M is a N+4 x N+4 matrix of the external forces

F is a vector of the spring forces and

reaction forces and moments

The spring forces lead directly to spring displacements and the details can be found in Norman (22).

The discrete support model gives results similar to the Winkler model for similar inputs. However, the discrete

model has the additional ability to represent non-uniform track.

Figure 7(B) compares the deflections from the two models for uniform modulus and a single applied load.

The continuous line represents the Winkler model and the boxes indicate the tie locations in the discrete model. The

track modulus used in the Winkler model was 20.7 MPa (3000 lbf/in./in.) and the corresponding tie stiffness was

10.5 MN/m (60000 lbf/in.). Track modulus is equated to tie stiffness by dividing by the tie spacing (ties spacing of

50.8 cm (20”)). A single point load of 157 kN (35750 lbf) was applied over the center tie. The deflection predicted

by both models is very similar with a maximum variation of 6.47%.

The clear advantage of the discrete support model is that it can represent non-uniform track. In Figure 7(C)

the stiffness of the 3rd tie from the left end has been decreased by 50% (to 5.25 MN/m or 30000 lbf/in.). In Figure

7(D), the stiffness of the 3rd tie has been increased by 100% (to 21.0 MN/m or 120000 lbf/in.).

The track deflection with a single soft tie (Figure 7(C)) is no longer symmetric about the loading point. The

rail is deflected more on the left side of the load where the soft tie is located. The maximum deflection of the rail

was also slightly increased (by approximately 0.1219 mm (0.0048 in.)). Figure 7(D) shows the rail deflection where

the stiffness of the 3rd tie has been doubled to 21.0 MN/m (120000 lbf/in.). The discrete model shows that the

deflection near the stiff tie and the maximum deflection have both decreased (by approximately 0.1829 mm (0.0072

in.)). The results from these examples show that 1) the two models give similar results for similar inputs, and 2) the

deflection curve can be affected by a single tie.


Lu et al. 11

Sheared Layer Model

A third solution of the BOEF model adds a shear layer to the uniform elastic rail foundation. In this model, known

as the Pasternak foundation, vertical displacement of one section of the elastic foundation will result in displacement

of neighboring sections of the elastic foundation (e.g. a mattress where the springs are tied together). This

distinction is most prevalent when the beam has low bending stiffness (i.e. low EI). Here, the supporting distributed

load, p(x) is given by:

d 2w
p( x) = −G p +upw (13)
dx 2

where up is a track modulus and Gp is a shear modulus. Substituting into Equation (1) gives the following governing

differential equation:

d 4w d 2w
EI − G p +upw = q (14)
dx 4 dx 2

The solution (from Kerr****) for a single applied load P acting at x=0, is

P β 2 −α x
w( x) = e [κ cos(κ x ) + α sin(κ x )],−∞ < x < ∞ (15)
2u p ακ

where:

up Gp (16)
β2 = ; α,κ = ± β 2 ±
4EI 4EI

The resulting relationship between applied load, P, and deflection, wo, is still linear as in the Winkler

Model, Figure 8. Here Gp=60GPa(8,702,400psi), I=3663cm4(88in4), E=206.8GPa (30,000,000psi) are used as

parameters. However, the effective stiffness of the Pasternak model is higher because more of the elastic foundation

is involved in producing reaction supporting pressure.

The difference between the Pasternak model and the Winkler model are more evident when either the beam

is not stiff (low EI) or the shear modulus is high. Figure 9 shows the correlation between the deflections of the two

models, for an identical beam under identical loads, with two shear modulus values. Again, as the shear modulus is

increased more of the elastic foundation produces supporting pressure resulting in both a stiffer track and an altered

shape. The very significant difficulty in using the model is in identifying an appropriate value of the shear modulus.
Lu et al. 12

Nonlinear Cubic Model

The limitation with all the previous models is that each uses some form of linear elastic behavior to represent the

supporting pressure. Field tests conducted by the ASCE-AREA Special Committee on Stresses in Railroad Track

(13) clearly showed that the vertical rail deflections were not linearly proportional to the wheel loads. An extensive

experimental study conducted by Zarembski and Choros (14) also clearly documented this nonlinear response.

Here, a new model is proposed that represents the relationship between vertical rail deflection and the rail

support distributed load as a cubic polynomial. To define this relationship the experimental results of Zarembski

and Choros (14) are plotted in Figure 10 along with a cubic polynomial curve fit. The polynomial fit is excellent

(R2=0.9987).

Using a cubic polynomial has several advantages. First, it clearly captures the behavior of real track

(Figure 10) in that it provides for low stiffness at low loads and higher stiffness at higher loads. Also, negative

displacement of the track (track lift) does not result in significant downward forces being applied to the rail. Unlike

the previous models, the cubic polynomial represents the fact that if the track rises slightly, the ballast does not pull

the track down.

Here, the supporting distributed load p(x) has a cubic relationship between p(x) and w(x):

p( x) = u1 w + u 3 w 3 (17)

Note, that symmetry about the applied load requires the second order term to vanish. Substituting into the BOEF

model gives the following differential equation.

d 4w
EI + u1 w + u 3 w 3 = q (18)
dx 4

Equation (19) is a nonlinear differential equation and a closed form analytical solution is not straightforward. One

analytical approximation based on the Cunningham’s method can be found in McVey (23). However, a numerical

solution for this Boundary Value Problem (BVP) can be obtained.

The BVP can be written in state space notation as:

w′ = func ( w, x) (19)
Lu et al. 13

⎡ w( x) ⎤
⎢ ⎥
∂ ⎢ w′( x) ⎥
w′ = = func( w, x) (20)
∂x ⎢ w′′( x) ⎥
⎢ ⎥
⎣ w′′′( x)⎦

Given equation (19) the BVP becomes:

⎡ w( x) ⎤ ⎡ w′( x) ⎤
⎢ ⎥ ⎢ w′′( x) ⎥
∂ ⎢ w′( x) ⎥ ⎢ ⎥
=⎢ w′′′( x ) ⎥ (21)
∂x ⎢ w′′( x) ⎥ ⎢
⎢ ⎥ − 1 u w( x) + u w 3 ( x) ⎥
( )
⎣ w′′′( x ) ⎦ ⎢⎣ EI 1 ⎥

3

As the name implies, the fourth order BVP described above requires the value of four boundary conditions, here:

w( x) | x =∞ = 0
w( x) | x =−∞ = 0
(22)
w′( x) | x =0 = 0
w( x) | x =0 = wo

Now, since the BVP can have more than one correct solution, an initial “guess” for the last boundary

condition that will cause the solution to converge to the expected solution. In this work, the initial guess is provided

by the Winkler model evaluated at x=0 and u=u3 given by:

Pβ ⎛ u ⎞4
w(0 ) = wo = − where : β = ⎜ 3 ⎟ (23)
2u 3 ⎜ 4 EI ⎟
⎝ ⎠

The mechanics of this problem also requires the solution be found subject to the additional constraint given

by the free body diagram in Figure 2 by:

∫ (u w + u w )dx = 2

3P (24)
1 3
0

The unique solution that satisfies all these constraints will give the rail deflection. Any number of

numerical techniques can be used to solve this well posed BVP. In this work the “bvp4c” function in Matlab (24)

was used.
Lu et al. 14

As the cubic model closely represents the deflection test data for the whole range of wheel loads, the

accuracy of the linear analysis depends on the magnitude of the test load.

Because the cubic spring is initially softer than the one in the Winkler model, the rail must deflect more

before the base can pick up the full load. This means that the distributed load will be spread over a wider span than it

is for the linear model as shown in Figure 11. Meanwhile, the deflection at the contact point for the cubic model is

slightly larger than the one for the Winkler model when the applied load is relatively large.

Track Modulus at Characteristic Load using the Cubic Model

Finally, the track modulus at characteristic load can be calculated:

u* =
∂p
=
(
∂ u1 w + u 3 w 3 ) = u1 + 3u 3 w 2 (25)
∂w P * ∂w P *
P*

This definition of track modulus is compared to the Winkler model for a given measurement of load of 151240 N

(34kips) and displacement of 0.254cm (0.1”) in Figure 12. In this Figure the load deflection curve is plotted from

the experimental data of Zarembski and Choros (14) shown in Figure 10.

It is clear that for single data points at higher loads the Winkler model will always underestimate the actual

track modulus (Figure 12). The Winkler model will also poorly represent changes in deflection with respect to

changes in load at these higher values. It is also clear from these data that any two choices of loads (as in the

Heavy-Light load definition of track modulus) will give a different value of track modulus.

CONCLUSIONS

Due to the widely accepted non-linearity of track response, the linear Winkler model obviously has its inadequacy.

Other models like the Pasternak model and the discrete model attempt to modify the Winkler model to develop

models that could more accurately describe an actual track foundation’s behavior under various applied loads, but

they are still based on the linear assumption. The heavy-light load method does provide a better approximation to the

nonlinear behavior, but there are still some discrepancies between the piecewise linear approximation and the real

continuous nonlinear track behavior.

The cubic model clearly captures the behavior of real track in that it provides for low stiffness at low loads and

higher stiffness at higher loads. It represents the real track structure under the whole range of wheel loads.
Lu et al. 15

Combined with the proposed definition of track modulus at characteristic load, the cubic model can sensitively

demonstrate the changes in deflection with respect to the changes in load at higher values, which the linear Winkler

model will poorly represent.

ACKNOWLEDGEMENTS

This work is supported under a grant from the Federal Railroad Administration. The authors would specifically like

to thank Mahmood Fateh and Gary Carr with FRA and William GeMeiner of the UPRR. We would also like to

thank BNSF and UPRR for track access and logistical support.

REFERENCES

[1] Cai, Z., Raymond, G. P., and Bathurst, R. J., 1994, “Estimate of Static Track Modulus Using Elastic Foundation

Models,” Transportation Research Record 1470, pp. 65-72.

[2] Selig, E. T., and Li, D., 1994, “Track Modulus: Its meaning and Factors Influencing It,” Transportation

Research Record 1470, pp. 47-54.

[3] Li, Dingqing, and Selig, Ernest T., 1994, “Resilient Modulus for Fine-Grained Subgrade Soils,” Journal of

Geotechnical Engineering, Vol. 120, No. 6, pp 939-957.

[4] Li, Dingqing, and Selig, Ernest T., 1998, “Method for Railroad Track Foundation Design I: Development,”

Journal of Geotechnical And Geoenvironmental Engineering, Vol. 68, No. 7-8, pp 457-470.

[5] Stewart, Henry E., 1985, “Measurement and Prediction of Vertical Track Modulus,” Transportation Research

Record 1022, pp. 65-71.

[6] Read, D., Chrismer, S., Ebersohn, W., and Selig, E., 1994, “Track Modulus Measurements at the Pueblo Soft

Subgrade Site,” Transportation Research Record 1470, pp. 55-64.

[7] Ebersohn, W., Trevizo, M. C., and Selig, E. T., 1993, “Effect of Low Track Modulus on Track Performance,”

International Heavy Haul Association, Proc. Of Fifth International Heavy Haul Conference pp. 379-388.

[8] Zarembski, A. M., and Palese, J., August 2003, “Transitions eliminate impact at crossings,” Railway Track &

Structures, pp. 28-30.

[9] Davis, D. D., Otter, D., Li, D., and Singh, S., December 2003, “Bridge approach performance in revenue

service,” Railway Track & Structures, pp. 18-20.


Lu et al. 16

[10] Carr, Gary A. and Greif, Robert, 2000. “Vertical Dynamic Response of Railroad Track Induced by High Speed

Trains,” Proc. of the ASME/IEEE Joint, pp. 135-151.

[11] Heelis, M. E., Collop, A. C., Chapman, D. N. and Krylov, V., 1999, “Predicting and measuring vertical track

displacements on soft subgrades,” Railway Engineering.

[12] Kerr, Arnold D. “On the Stress Analysis of Rails and Ties,” Proceedings American Railway Engineering

Association, 1976, Vol. 78, pp 19-43.

[13] ASCE-AREA Special Committee on Stresses in Railroad Track, Bulletin of the AREA, First Progress Report

Vol 19,1918, Second Progress Report Vol 21,1920.

[14] Zarembski, Allan M. and Choros, John. “On the measurement and calculation of vertical track modulus,”

Proceedings American Railway Engineering Association, 1980, Vol. 81, pp. 156-173.

[15] Kerr, A. D. and Shenton, H. W.,1986, “Railroad Track Analyses and Determination of Parameteres,” Journal of

Engineering Mechanics, Vol. 112, No.11, pp1117-1134.

[16] Ebersohn, W., and Selig, E. T., 1994, “Track Modulus on a Heavy Haul Line,” Transportation Research Record

1470, pp. 73-83.

[17] Kerr, A.D., 2003, “Fundamentals of Railway Track Engineering,” Simmons-Boardman Books, Inc., Omaha,

pp. 137-152.

[18] Arnold, R., Lu, S., et al, 2006, “Measurement of Vertical Track Modulus from a Moving Railcar,” AREMA

Conference Proceedings 2006.

[19] Raymond, G. P., 1985, “Analysis of Track Support and Determination of Track Modulus,” Transportation

Research Record 1022, pp. 80-90.

[20] Meyer, Marcus B. 2002, “Measurement of Railroad Track Modulus on a Fast Moving Railcar,” University of

Nebraska – Lincoln. May 2002.

[21] Boresi, Arthur P. and Schmidt, Richard J. 2003. Advanced Mechanics of Materials 6th Edition. John Wiley &

Sons, New York, NY: Chap. 5,10.

[22] Norman, Christopher D. 2004, “Measurment of Track Modulus from a Moving Railcar,” Masters Thesis,

University of Nebraska – Lincoln. August 2004.


Lu et al. 17

[23] McVey, Brian. 2006, “A Nonlinear Approach to Measurementt of Vertical Track Deflection from a Moving

Railcar,” Masters Thesis, University of Nebraska – Lincoln. May 2006.

[24] Kierzenka J. and Shampine L. F., 2001, “A BVP Solver based on Residual Control and the MATLAB PSE,”

ACM TOMS, Vol. 27, No. 3, pp. 299-316.


Lu et al. 18

Figure 1: Free Body Diagram of the Rail


Figure 2: Boundary Condition of the Rail
Figure 3: Deflection of Track Under Three Loads
Figure 4: Various Representations of Track Modulus
Figure 5: Relative Rail Displacement Under a Railcar
Figure 6: Discrete Model and Free Body Diagram
Figure 7: Comparison of Winkler and Discrete Models
Figure 8: Stiffness of Winkler and Pasternak Models
Figure 9: Comparison of Winkler and Pasternak Models
Figure 10: Experimental Data and Curve Fitting
Figure 11: Comparison of Cubic and Winkler Models
Figure 12: Modulus Calculations in Winkler and Cubic Model
LIST OF FIGURES

Figure 1: Free Body Diagram of the Rail 


Figure 2: Boundary Condition of the Rail 
Figure 3: Deflection of Track Under Three Loads 
Figure 4: Various Representations of Track Modulus 
Figure 5: Relative Rail Displacement Under a Railcar 
Figure 6: Discrete Model and Free Body Diagram 
Figure 7: Comparison of Winkler and Discrete Models 
Figure 8: Stiffness of Winkler and Pasternak Models 
Figure 9: Comparison of Winkler and Pasternak Models 
Figure 10: Experimental Data and Curve Fitting 
Figure 11: Comparison of Cubic and Winkler Models 
Figure 12: Modulus Calculations in Winkler and Cubic Model 

View publication stats

Das könnte Ihnen auch gefallen