Hou Dissertation 2020
Hou Dissertation 2020
BRIDGE APPROACHES
BY
WENTING HOU
DISSERTATION
Urbana, Illinois
Doctoral Committee:
that can analytically tackle the difficult problem of a railway track transition zone and
effectively simulate the dynamic response of a bridge approach subjected to moving train
loads. The scope of the research encompassed statistical field data analysis, analytical
track model to provide improved design and maintenance practices. The objective was to
better understand governing mechanisms of bridge approach problems that occur near
associated with differential settlement, hanging ties, impact loads and various lack of
support conditions, select mitigation methods for existing deficiencies, and as a result,
The first task in this PhD study involved identification of problematic track bridge
interpretation along the US Amtrak Northeast Corridor lines near Chester, Pennsylvania.
monitor individual layer deformations of track substructure layers. Besides MDDs, strain
gauges were also mounted on the rail at the instrumented sites to measure vertical wheel
loads applied during the passage of a train. Statistical analyses were conducted on the
collected field data to quantify track transient response and performance trends at the
ii
Field collected data indicated one significant problem occurring at the instrumented
bridge approaches, also known as the “hanging crosstie” problem, which is caused by
several sequential crossties near the bridge abutment experiencing lack-of-support. These
crossties with gaps formed underneath undergo oscillatory motion as the dynamic loading
from moving wheels push and pull these crossties through rail deformations. As a result,
the nonuniform support conditions of the track substructure worsen drastically and can
result in extremely high deformations including heave condition due to train passage.
solution scheme for both open track (regular track) and near bridge (with severe “hanging
crosstie” problem) locations. The simulation results from the developed ballasted train-
track-bridge model matched well with the deformation data collected from the field
instrumentation to properly validate the analytical solution scheme. Then, using the field
validated track model, various mitigation methods, including changing rail pad stiffness,
ballast stiffness, subgrade stiffness, and crosstie spacing, applicable to track bridge
approaches were studied and discussed. Furthermore, details on two successfully applied
mitigation methods, namely, (i) converting an open deck bridge to ballasted deck bridge
and (ii) installation of a new track panel consisting of approximately 30 concrete ties with
Under Tie Pads (UTPs) at a bridge entrance on Amtrak’s Northeast Corridor high speed
iii
ACKNOWLEDGMENTS
First and foremost, I would like to sincerely thank my advisor, Professor Erol
Tutumluer, for his continuous support and encouragement. His enthusiasm for research and
innovation has always been a source of inspiration for me. I am also grateful for the
numerous professional and academic opportunities he has provided me over the past six
This dissertation would not have been possible without the many wonderful people
with whom I had the pleasure to collaborate. I am grateful to Drs. Hasan Ozer and Issam
Boler for the great time working together on railroad track transition project. Besides, I am
indebted to Bin Feng and Wei Li for their hard work when we collaborated on the railroad
crosstie support condition project. I gratefully thank Drs. Christopher Barkan, Billie F.
Spencer Jr., Hai Huang, Debakanta Mishra for taking part in my Ph.D. thesis committee
I would like to thank my friends, Yubing Liang and He Wang, who have provided
me with constant support and brought tremendous joy to my graduate life. My special
thanks are to my colleagues Priyanka Sarker and Wenjing Li, for their continuous kindness,
Last but not least, my deepest love and gratitude are for my family, my father
Kaifeng Hou, my mother Jun Wang, and my husband Tuo Zhao, for their unconditional
can describe how grateful I am for everything they have done for me. It is with great
iv
TABLE OF CONTENTS
v
CHAPTER 1: INTRODUCTION
1.1 Introduction
While most railroad track structures will not experience sudden changes over a
short distance, there still are cases where the adjacent track structures exhibit totally
which results in different characteristics noted as discontinuities along the track structure.
Due to the sudden change in track stiffness, the “stiff” side of a track transition, e.g. bridge
deck, undergoes lower deformations due to train loading, compared to the “less stiff” side,
e.g. open track side. Such differential movement caused geometry problems at railway
transitions have been well recognized, and they present a significant challenge to
railroaders as far as maintenance of track profile is concerned (Li and Davis, 2005;
Design standards including European Rail Research Institute (ERRI), UIC code
719, Design Standards for Railway Structures and Commentary, and AREMA all involve
the designs of a transition zone (Muramoto, 2013). Track transition zones require frequent
track (Kerr and Maroney, 1993). The Association of American Railroads (AAR) reported
transitions (Sasaoka and Davis, 2005). Similarly, more than USD 110 million was spent
annually on transition zones in Europe by 1999 (Hyslip et al. 2009). According to the AAR
data center, America's freight railroads spend $25 billion a year on average from 2006 to
1
2018 on capital expenditures and maintenance expenses related to locomotives, freight cars,
The most common transition problems are often encountered at bridge approaches,
with the approach track on either side of the bridge abutment being much less stiff
compared to the bridge deck often supported by deep foundations. Figure 1.1 illustrates a
typical structural view of the transition zone at track bridge approach. The “embankment
side” on the right-hand side represents regular ballasted track structure. And the “bridge
side” on the left-hand side represents an open deck bridge where there is no ballast layer.
As shown in the figure, the abutment at bridge side is often non-deformable and modulus
properties of the layers on both sides could be significantly different. In accordance, the
tracks on a bridge deck undergo significantly lower deformations under loading compared
to the approach tracks. This sudden change in track deformation behavior at the transition
point may result in extreme dynamic loading conditions and may ultimately lead to rapid
deterioration of the track and bridge structural components (Mishra et al. 2012).
2
One of the most influential results of the differential movement occurring at bridge
crossties near the bridge abutment experiencing lack-of-support (Kerr and Maroney, 1993;
Li and Davis, 2005; Muramoto et al. 2012; Banimahd et al. 2012; Sanudo et al. 2016).
These crossties with gaps formed underneath undergo oscillatory motion as the dynamic
loading from moving wheels push and pull these crossties through rail deformations. As a
result, the non-uniform support conditions of the track substructure worsen drastically and
can result in extremely high deformations including heave condition due to train passage.
Figure 1.2 illustrates an example of unequal tie-ballast gap modeled by Stark et al. (2015)
bridge approaches, improve safety, and ensure passenger comfort, the mechanism of
dynamic responses at railroad bridge approach must be properly identified. This requires
the combined application of field instrumentation for measuring dynamic track responses
3
1.2 Problem Statement
have been found in the vicinity of bridge abutments (Sanudo et al., 2016). The working
conditions of bridge approaches under dynamic loading have great impact on the rider
comfort and operational safety. Track geometry also suffers from accelerated deterioration
at bridge approaches.
essential to consider the coupled effect of moving train, open track, hangint crossties and
bridge deck together. To mend the gap in literature, this dissertation focuses on developing
a model to ideally consider discretely supported track for both the open track side and the
bridge side. The open track side includes multiple layers underneath the rail including
crosstie, ballast, and subgrade. The bridge side adequately models the effect of the stiff
open deck bridge. In addition, the hanging crosstie problem that often occurs at the bridge
substructure when there is a gap between crosstie and ballast particles. The developed
model can be utilized to simulate responses at bridge approaches under real train moving
potentially beneficial to track behavior under loading. However, it is often difficult and
expensive to try different combinations of mitigation methods in the field. Therefore, the
methods.
4
1.3 Objectives
In this PhD study, field monitoring data collected from railroad track bridge
approaches are analyzed to determine track dynamic responses and deformation behavior.
A dynamic track model aimed at simulating track behavior at bridge approaches under
moving load is developed. Both open track and bridge deck are included in the simulation
between crosstie and ballast contact forces and relative displacements are considered in the
dynamic track model. The detailed objectives of this research study are listed as follows:
(1) Develop a train-track-bridge model at the bridge approach that can realistically
predict responses under dynamic train loads, for example, due to the observed
difficult cases of hanging crossties and lack of support conditions, and validate the
(2) Use the validated model as an analytical solution scheme to study the effects of
changing crosstie spacing and substructure layer properties, converting open deck
science and engineering with the development of a dynamic track model at transition zones,
specifically ballasted railway bridge approaches, that will eventually help engineers to
better simulate dynamic response behavior of track structures under moving loads. The
validated track model is expected to serve as an analysis tool to provide improved design
5
1.4 Methodology and Scope
As part of recent track transition zone research efforts at UIUC, major field
instrumentation tasks were undertaken at several railroad bridge approaches from July to
August of 2012. Multi-Depth Deflectometers (MDDs) and rail strain gauges were installed
on three Northeast Corridor bridge approaches of Amtrak passenger lines near Chester,
Pennsylvania, USA. Transient deformation and wheel loading data from both passenger
and freight trains were recorded at the instrumented sites at different time intervals.
As part of the study scope, the collected field data are first reduced and analyzed
using statistical approaches including 95% confidence interval. Next, individual layer
transient displacements from the MDD data are used to the backcalculate the instrumented
track substructure layer moduli using the field measured wheel loads in order to evaluate
locations (open track vs. near bridge). This step employed GEOTRACK, a well-known
followed to simulate the behavior of railroad bridge approaches under dynamic loading.
consider the realistic coupled conditions of the open track and the bridge deck sides of the
bridge approach. The derived model is then solved with Newmark integration method and
implemented in MATLAB for numerical results. This step is intended to simulate open
track behavior at bridge approaches under dynamic single moving wheel load and compare
6
the predicted results with field measured responses on the open track side of the bridge
consider nonlinear stiffness, which represents unsupported crossties near the bridge
numerical simulation method Newmark-𝛽 Integration Method for numerical results. Given
that the focus of this study is the track substructure of bridge approaches, the bridge is
then validated with field measured responses collected from the bridge approaches (near
parameters, and structural properties of the field conditions are used to accomplish model
validation. The validated model can be used to conduct sensitivity analysis on train speed
Finally, several mitigation methods are studied using the validated train-track-
bridge model. The proposed model is employed to study vertical transient deformations of
track structure under moving loads with the help of possible mitigation methods such as
changing crosstie spacing and substructure properties, converting open deck bridge to
ballast deck bridge, and installation of crossties with under tie pads (UTPs) at transition
zones. Results from the model simulations are used to propose practical solutions for bridge
railroad track bridge approaches and utilize the validated train-track-bridge model to aid in
selecting mitigation methods for remedying existing problematic bridge approaches and
7
developing improved transition zones. For example, by identifying the effects of track
properties, vehicle speed, and hanging crossties on the performance trends of both open
track and near bridge locations, the designs of transition zones could be improved, which
would result in adequate track modulus and support with little differential movement.
This thesis consists of seven chapters. The remaining chapters are organized as
follows:
Amtrak’s Northeast Corridor lines. Next exploratory data analyses on the collected field
instrumentation data are presented. Chapter 3 also presents inferential statistical data
moduli using field MDD deformations and wheel load data are presented. These calculated
moduli are used as a base line for estimating layer stiffness characteristics in the developed
well as numerical and analytical solution schemes of the developed ballasted train-track-
bridge model for both the linear model at regular embankment track structure and the
nonlinear model at near bridge location. The proposed nonlinear track model adequately
8
Chapter 5 presents the numerical simulation results for two case studies utilizing
the established ballasted train-track-bridge model. The developed track model simulates an
Amtrak passenger train moving on top of track bridge approaches. Results for both open
track location (linear model) and near bridge location (nonlinear model) are compared with
field instrumentation data for validation purposes. The validated model is then used to
generate more key indexes of track dynamic response. In addition, the validated model is
also utilized to study the train speed and train loading or weight effects on track dynamic
responses.
different mitigation methods for railroad track bridge approaches including changing
crosstie spacing and substructure properties, converting open deck bridge to ballasted deck
bridge, and installation of under tie pads on several crossties at the bridge approach
transition zones.
Finally, Chapter 7 summarizes the research effort, findings and conclusions. Future
9
CHAPTER 2: LITERATURE REVIEW
This chapter provides an introduction to railway track structures and terms, types
and performance records of track transitions as bridge approaches. Recent work on bridge
approaches is then reviewed. Mitigation efforts are also summarized in this chapter.
and energy efficient ways to transport both people and freight. Figure 2.1, for example,
shows that intercity rail is the most efficient mode for the fuel economy among the 9
There are two major types of railroads from the perspective of structural layered
systems. One is called ballasted track, which is a traditional type of track structure,
consisting of rail, crossties, and ballast layer. The other type of track system is non-
10
ballasted track, which substitutes ballast with concrete slabs. The two types of track
systems both have their own advantages and are used widely worldwide. Ballasted track is
a traditional railroad track system that is usually used in the normal speed passenger and
freight lines, while ballastless track has been more and more adopted in high-speed lines.
Ballasted track requires a lower initial construction cost but higher maintenance cost while
ballastless track requires a higher construction cost but lower maintenance cost.
freight railroad network is nearly 140,000 miles. Its general components include ballast,
crossties, tie plates and rail, as shown in Figure 2.2. Note that crossties are referred as
“sleepers” in Figure 2.2. Ballast is used as the granular layer to provide ease in
constructability, stability, load distribution, and drainage ability to the main track
component of a ballasted railway system for several decades. The ballast layer is mostly
manufactured from local materials, preferably crushed stone. Ballast is installed under
crossties for transferring from loads from passing trains. It is required to maintain the track
in both good horizontal and vertical alignment. Note that track geometry deterioration can
proper drainage to the track structure. Kaewunruen and Tang (2019) noted that to fulfill
1. Durable to absorb the loads from crossties and transfer the loads to the subgrade
11
4. Suitable particle size to allow packing and load transmitting, and sufficient void
loads and transmit them to the ballast to maintain uniform rail gauge for safety. They can
also improve stiffness and stability of track structure by providing lateral track resistance.
Crossties are usually made of timber, steel, concrete and plastic or composite materials,
while concrete is the most adopted material in high speed routes. The materials can largely
Tie plates are used to support rails and fix the rail fastening system to provide a
smooth and uniform bearing surface for rail and crossties, prevent the rail from cutting into
Rail is mostly made of steel, which can help guide the train, withstand the wheel
loads and transmit the loads to crossties. Rail must provide a smooth and continuous rolling
12
2.2 Track Transition: Definition, Types, and Performance Records
While most railroad track substructures do not experience sudden changes in the
operating areas, there are still many cases where intentionally or passively, the track
substructures across a very short distance would exhibit tremendous differences. Examples
of railway track transitions include tunnels, special track work, highway/rail at-grade
In the U.S railroad network system, there are over 100,000 bridges, generating great
discontinuities along the track. A track transition zone is defined as a certain section of
track which results in different characteristics noted due to such discontinuities along the
track structure. Due to the sudden change in track stiffness, the “stiff” side of a track
transition, e.g. bridge deck, undergoes lower deformations under loading, compared to the
Many problematic issues arise at railroad track transition zones. One of the most
significant problems is differential settlements, which occur between two sides of the
transition point, and has been a global problem for both track maintenance personnel and
of approximately USD 200 million to maintain track transitions (Sasaoka et al. 2005,
Hyslip et al. 2009), whereas more than USD 110 million was spent annually on transition
of track profile is concerned (Woodward et al. 2007, Banimahd et al. 2012). Due to the
sudden change in track stiffness, the “stiff” side of a track transition undergoes lower
deformations under loading, compared to the “less stiff” side. This differential movement
13
often results in the formation of a “bump” in the track profile. Differences in track system
stiffness and/or damping characteristics, settlement of the ballast layer due to degradation
and/or fouling, and settlement of the subgrade and/or fill layers are some of the factors
requires the combined application of field instrumentation with analytical and numerical
track modeling.
Bridge approaches, as one of the most significant track transitions, have been
investigated by many researchers. The “bump” and “dip” formations at the end of a railroad
bridge were illustrated by Nicks (2009) as shown in Figure 2.3. Due to drastic differences
in substructure and loading conditions, the tracks on a bridge deck undergo significantly
lower deformations due to train loading compared to the approach tracks. This sudden
change in track deformation behavior at the transition point results in extreme dynamic
loading conditions and ultimately leads to rapid deterioration of the track and bridge
development of a “bump” usually within 15 m from the bridge end (Plotkin and Davis
2008).
indicated that approximately 50% of bridge approaches developed a low approach, usually
6 to 102 mm in depth and 1.2 to 15.2 m in length, that adversely affected ride quality
(Briaud et al. 2006). Development of sudden dips adjacent to the bridge deck increases the
14
dynamic impact loads significantly. Koch and Hudacsek (2017) reported vertical dynamic
loads about twice the static wheel load level for coal gondolas at track transitions. Read
and Li (2006) concluded that the bump problem at track transitions was more significant
as a train moves from a high-stiffness track to a low-stiffness track. According to Read and
Li (2006), the problem of differential movement was more critical at the exit end of a bridge,
whereas the sudden track stiffness increase at the bridge entrance led to rail surface fatigue,
Figure 2.3 The Bump (left) and Dip (right) at End of Railway Bridge (Nicks, 2009).
for high-speed rail infrastructure as the “bump” is accentuated at high speeds. The issue is
even more critical for shared corridors carrying both freight and high-speed lines.
Transitions along shared corridors need to be maintained to satisfy the high ride quality
requirements associated with high-speed trains. Additionally, these transitions also need to
withstand the heavy loads imposed by slow-moving freight trains without undergoing
excessive deformations. With the current impetus for development of high-speed lines in
the U.S and the challenges associated with shared corridors for operation of passenger
trains at increased speeds, preventing and mitigating the problem of differential movement
at bridge approaches and other track transitions has become more significant.
15
One of the most influential results of the differential movement occurring at bridge
approaches is the “hanging tie” problem, which is caused by several sequential crossties
near the bridge abutment experiencing lack-of-support (Kerr and Maroney, 1993; Li and
Davis, 2005; Muramoto et al. 2012; Banimahd et al. 2012; Sanudo et al. 2016). These
crossties with gaps formed underneath undergo oscillatory motion as the dynamic loading
from moving wheels push and pull these ties through rail deformations. As a result, the
nonuniform support conditions of the track substructure worsen drastically and can result
The root causes and major influential factors affecting bridge approach problems
are not fully agreed upon by researchers. From the investigation of four bridge approaches
with concrete ballast-deck bridges and concrete ties, Li and Davis (2005) reported
inadequate ballast and subballast layer performance to be the primary cause of track
geometry degradation. Using settlement rods installed in the test sections, they observed
deterioration for a site with cement-stabilized backfill. On the other hand, Selig and Li
(1994) identified subgrade stiffness to be the most influential parameter affecting the
moduli of ballasted tracks. As track transition problems are often related to the stiffness of
the approach trackbed, this would indicate that the subgrade layer plays the most significant
role in governing the differential movement at track transitions. Yet, recent field
instrumentation data to be discussed in detail in this dissertation clearly identify the ballast
layer as the primary contributor to the differential movement problem (Mishra et al. 2012,
Tutumluer et al. 2012, and Mishra et al. 2014). As the ballast “densifies” under repeated
16
train loading, related ballast movement and degradation often lead to recurrent ballast
differential movement at track transitions, a general consensus exists regarding the list of
plausible mechanisms. Sasaoka and Davis (2005) attributed track transition problems to
three primary factors: (1) differential settlement, (2) differences in stiffness characteristics,
and (3) discrepancies in track damping properties between adjacent sections. Similarly, Li
and Davis (2005) listed (1) track stiffness change, (2) ballast settlement, and (3)
geotechnical issues as the major causes of bridge approach problems. Nicks (2009) listed
development at railway bridge approaches: (1) differential track modulus, (2) quality of
approach fill, (3) impact loads, (4) ballast material, (5) drainage, (6) damping, (7) abutment
type, (8) bridge joint, (9) traffic considerations, and (10) quality of construction. Note that
influencing the differential movement and other track deterioration problems at transitions,
Plotkin and Davis (2008) used five different analysis methods to conclude that stiffness
differences did not play an important role as far as track behavior and ride quality at track
structural behavior under moving trains. Several attempts have been made in the past to
characterize “bump” development at track transitions. For instance, strain gauges (see
17
Figure 2.4) are widely used for measuring wheel loads on the rail. Namura and Suzuki
(2007) used axle box acceleration data to evaluate the wheel load on vehicles and placed
strain gauges on rail webs to measure the wheel loads on the rail. The field measured data
were compared with analytical model results to validate effectiveness of the modeling
approach. Mitigation methods were then proposed for minimizing the track geometry
degradation at track transitions. Sakurai et al. (2013) adopted a set of instruments for
track were pulled down against ballast by tie rods fixed to the anchors laid under trackbed.
Strain gauges were placed on tie rod to measure reactions (Sakurai et al. 2013). Similarly,
Hayano et al. (2013) investigated the effect of ballast thickness and tamping repair
settlement of footing under cyclic loading. Computer vision techniques, such as 3D images,
were also used by researchers to measure soil mass density of ballast material (Hayano et
al. 2013). They introduced the method of using 3D digital camera to obtain stereo pair
images of excavated holes and estimate volume of the holes accordingly. The results
18
Figure 2.4 Dual-Element Shear Strain Gauges Installed at the Rail Neutral Axis for
Wheel Load and Tie Reaction Measurements.
As a quite effective method to study dynamic track response due to train loading,
Multidepth Deflectometers (MDDs), which are the subject of this research effort, have
been used to measure track substructure layer deformations. The MDD technology was
first developed in South Africa in the early 1980s to measure individual layer deformations
in highway pavements (Scullion et al., 1989). MDDs typically consist of up to six linear
small diameter (typically 45 mm diameter as also used in the current study) hole to measure
the displacements of individual substructure layers with respect to a fixed anchor buried
deep in the ground (De Beer et al., 1989). Although several studies in the U.S. have
their use in railroad applications has been limited. Two previous studies in the U.S.
(Sussmann and Selig 1998, Bilow and Li 2005) used MDDs to monitor the deformations
19
in railway track substructure layers. It is noteworthy that the use of MDDs to monitor
railway track performance has been extensively pursued in South Africa (Grabe and Shaw
2010, Priest et al. 2010, Vorster and Grabe 2013). The vertical deformations are measured
from a reference head near the top surface to predetermined depths in a borehole with the
overall vertical deformation referenced to a fixed anchor in the subgrade (see Figure 2.5).
Coelho et al. (2011) studied typical transition zone in the Netherlands. In their research,
accelerations and velocities of the track, soil, and approach slabs in response to passenger
trains were measured for the calculation of displacements. In addition, track settlements
(a) (b)
20
2.5 Analytical and Numerical Modeling of Railroad Track Transitions
There have been various mathematical models developed to interpret and predict
the dynamic response of railroad track. Early models were one- or two-dimensional
composed of a beam as the rail on a Winkler foundation under moving force as wheel load.
Mise and Kunii (1956) developed a theoretical framework for the vibrations of a flexible
beam supporting a moving locomotive. In their simplified model, only a beam under a
moving force was considered. Kalker (1996) introduced discretely supported rails under a
travelling vertical point load to better represent the irregular discrete support of the
crossties. Also, closed-form solutions were derived for the dynamic response of simple
beam under high-speed trains (Yang et al. 1997). Huang et al. (2009) developed the
analytical solution of a railroad track model under moving loads with asphalt trackbed used
underneath ballast layers (see Figure 2.6). Basu and Rao (2013) studied the steady state
21
A discretely supported rail system with coupling of soil’s shear resistance can
adequately represent the track and the substructure in the field. However, these models do
not include the mass of train and the interacting force between wheel and track or
irregularity of track profile. The work of Lee (1998) introduced dynamic response of a
Timoshenko beam on Winkler foundation subjected to moving mass load and showed the
limitation of neglecting inertial effects of the mass when only equivalent moving force was
considered. Furthermore, Lei and Noda (2002) developed a computational model for the
vehicle and track coupling system where the system was divided into two parts. Upper
structure is a whole locomotive with two layers of spring and damping system in which
vehicle and bogie are involved. The lower structure is a railway track where rails are
irregularity of the track vertical profile is also simulated in their work. Zhai and Sun (1994)
also investigated a model for vertical interaction between vehicle and track. Vehicle
on the track with a constant velocity, and the track substructure as an infinite Euler beam
supported on a discrete continuous elastic foundation consisting of the three layers of rail,
crosstie, and ballast. Lei and Rose (2008) analyzed track vibration by Fourier transform
Note that the above analytical models do not include the three-dimensional (3D)
dynamic wave field generated in the ground due to the passage of a train. Therefore, they
are restricted to the cases where the velocity of the train is much smaller than a critical
velocity (Krylov, 1995). Krylov (1995) studied ground vibrations generated by superfast
trains theoretically. He found that superfast railway trains moving with speeds approaching
22
or exceeding the Rayleigh wave velocity in the ground could cause very large increases in
ground vibration levels when compared to those generated by conventional trains. The
most typical Rayleigh wave velocity values for soils are 250-500m/s. Holm et al. (2002)
confirmed that “critical speed” is the speed at which trains induce a large resonant response
in the track with soft ground conditions resulting in excessive vertical vibrations. Based on
these theories, three-dimensional models on the soil considering ground vibrations were
proposed. Bian et al. (2008) conducted simulation of high-speed train induced ground
vibrations using a 2.5D finite element method. The train induced track and ground
vibrations have been studied using analytical-numerical combined method. Later on, a fully
coupled 3D train track soil model was also introduced by Huang et al. (2014) where three-
dimensional soil parts are included. Li et al. (2016) applied a 3D finite element model with
a cycle domain constitutive model for track settlement simulation. This study adopted a
continuum modelling of ballast settlement. The CDM could calculate effective crosstie
loads, while the contact pressure between crosstie and ballast was also calculated for every
vehicle passage. The method considered that the time of loading applying on adjacent
crossties was not the same. A full scale test track was built in Spain to validate the model
and calibrate chosen parameters. The results showed this model could predict the track
Most of the developed analytical models are linear in nature and therefore limited
when nonlinear aspects become significant, for example, when there is a hanging tie
problem. To avoid the limitation, Tanabe et al. (2003) built a 3D numerical analysis model
where nonlinear springs and dampers were employed. Nielsen and Oscarsson (2004)
23
expanded their previous model to account for state-dependent track properties, separated
that are dependent on the time-variant state of the different track components due to the
dynamic loading from a moving train model. Ling et al. (2014) developed a 3D dynamic
model of a nonlinear HSR coupled with a flexible ballast track. In this model, the vehicle
was modeled as a 42 degrees of freedom multi-body system, the track was modeled as a
traditional three-layer discrete elastic support model, the rails were assumed to be
Timoshenko beans supported by discrete ties, which were modeled as Euler beam and the
ballast bed was treated as equivalent rigid ballast body. After validation using numerical
comparison with software SIMPACK, this model was proved that it considered the mutual
influence of the adjacent vehicles and tracks, and the fast dynamics calculation on a long
train running on an infinitely long flexible track was possible to be carried out.
There are very few analytical solutions for track transition problem, in which case
mathematical model to study the influence of spatially varying track stiffness on the
performance of the wheel and the track at low frequencies. Biondi et al. (2005) investigated
the 2D dynamic interaction among a running train, a track structure and a supporting bridge
where both rails and bridge were modeled as Bernoulli Euler beams. The work of Coelho
et al. (2009) showed the importance of hanging crossties in the transition zone due to large
settlement of embankment and the higher dynamic impact forces induced by passing trains
on these areas. Zhai et al. (2013) established a model for the analysis of train-track-bridge
dynamic interactions which also accounts for the continuity problem. Varandas et al. (2011)
and Varandas (2013) confirmed the main cause for the higher displacement amplitudes
24
registered on the transition zone was the existence of a group of consecutive hanging
crossties. A nonlinear model (see Figure 2.7) was presented for the interaction between
force and displacement of springs and also took into account the unloaded position of track
and ballast (including hanging distances). The model can simulate when the wheels moving
onto the trough existing after the culvert as a lever effect that will lift the rail/crossties
system before the culvert. Yang et al. (2013) studied the dynamic responses of bridge-
numerical model (see Figure 2.8) of vehicle-track-subgrade coupling system based on the
25
Figure 2.8 Vertical dynamic analysis model of vehicle-track-subgrade coupled system
(Yang et al., 2013).
Indraratna et al. (2019) pointed out that a mathematical model based on beams on
elastic foundations was applied extensively in practice. The approach modeled the rail as
al. (2017) showed that both Euler-Bernoulli beam and Timoshenko beam gave almost the
same results.
testing are not commonly possible or the choice for studying behavior of a railway track
system due to long time commitments and often large efforts/costs associated. Yet
analytical models, such as the methods discussed in the previous section, are not always
accurate for the specific details and different geometries, which makes numerical modeling
a viable alternative for studying the track system under train loading. GEOTRACK (Chang
et al. 1980) is a layered elastic system program which has been validated and widely used
26
for track structural analysis. In GEOTRACK, stresses and deformations are calculated as
a function of multi-axle loads, properties of rails and ties, properties of substructure layers,
and geometry details of ties and underlying layers (Li and Selig, 1998). KENTRACK
(Huang et al. 1984) is a finite element-based trackbed structural design program that can
be utilized to analyze responses of granular ballast trackbed as well as asphalt trackbed and
slab track. Of course, these models also come with certain assumptions and several
limitations such that in some situations they may not be appropriate. As a limitation, for
instance, they do not consider dynamic response behavior of the track system.
To study the time-dependent behavior of track under dynamic loading, often the
Finite Element Method (FEM) has been adopted and used by researchers as the numerical
modeling method of choice (Feng 2011, Smith et al. 2006). Feng (2011) used FEM to
simulate different track systems with various analytical formulations and complexities,
such as beam on discrete support model, track with ballast mass model, and rail on crosstie
on continuum model. For individual cases, the displacements of trackbed were evaluated.
Numerical solutions indicated that railpad stiffness considered in the model had a major
effect on the resonance frequency. Wang and Valeri (2018) also applied FEM method for
transition zones. In the analysis, subgrade, ballast and crossties were modeled as 3D elastic
elements, and the bridge was modeled as a two layer structure with support layer and
concrete slab, and all of these elements had linear isotropic elastic material properties.
Since the model validation approved to provide accurate results, the FEM model was
presents that the settled area would gradually propagate from locations close to the bridge
27
to farther locations. Wang and Valeri (2019) improved the numerical model by
incorporating rails and crossties as beam elements, fasteners as spring elements, ballast as
3D solid elements, vehicles as mass-spring system. To obtain more accurate results and
detailed ballast stress analysis, they modeled rails as Hughes-Liu beam elements, crossties
as 3D elastic bodies, and these elements had the linear isotropic elastic material properties.
Meanwhile, the contact between ballast and crossties was modeled as nonlinear with
employing the penalty method and considering the ballast stress determining the permanent
settlement in ballast.
At track transitions where hanging ties are common (see Figure 2.9), tracks
However, only a few numerical models include these features (Smith et al. 2006,
Kaewunruen et al. 2011, Giner et al. 2012, Wang et al. 2015). Smith et al. (2006) conducted
a parametric study using a finite difference program FLAC3D, considering the effects of
velocity and stiffness of substructure materials. The limitation of this model is that the train
load was not distributed as in reality through the rail and ties. Kaewunruen et al. (2011)
used a nonlinear solver in STRAND7 to conduct numerical FEM simulations of track and
focused on hanging tie problem due to lack of support or excessive track settlement. The
model was able to evaluate the effect of voids on the dynamic responses of concrete crosstie.
The negative dynamic bending moments were enlarged by hanging ties and thus could
cause ballast to wear out sooner. Yin and Wei (2013) developed an FEM of railway bridge
transition zone using ABAQUS commercial software. The model included vehicle
modeling, bridge modeling, as well as open track modeling. Track irregularity was also
included in the model. More recently, a dynamic FEM model using explicit integration of
28
the track transition zone was developed (Wang et al., 2015) to model transition zones with
differential settlement or hanging ties. This model was validated through field
measurements on a ballast track. Kouroussis et al. (2015) studied the influence of rolling
the rail unevenness and other irregularity on the rail/wheel surface, this study first
constructed a numerical model with a flexible track on which the vehicles traverse at a
constant speed, then using a two-step approach to simulate ground wave propagation. In
this model, multibody formalism was used for vehicle modeling, while 2D approach was
preferred, the rail was modeled by a Euler-Bernoulli beam, the rail masses were modeled
using Timoshenko formulation, and the vertical wheel/rail contact forces calculation used
Hertzian theory. When Kouroussis et al. (2017) studied railway vibration, a combined
model. The additional experimental setup was applied to help solve the complexities in
numerical modeling caused by wheel impact local defects. The 3D dynamic model
established by Koch and Hudacsek (2017) used the PLAXIS 3D Dynamic package. Besides
modeling the rail as beam element, applying standard fixities and absorbent boundaries,
and setting certain values of material properties of crossties and rails, PLAXIS 3D used a
time-force signal to define the dynamic loads. There was a multiplier for each single
dynamic point load. But the model with different structure types and soil conditions should
on the long-term track behavior. Wilk (2017) also developed a numerical model for LS-
DYNA analyses that simulates a train truck entering an open deck bridge. The tie-ballast
29
gaps are considered in this model. Simulation results show that the differential stiffness
between bridge deck and transition zone leads to 20% increase in dynamic wheel load.
Also, evenly increasing the ballast settlement in the transition zone produced tie-ballast
gaps and increased tie loads by over 100%; while unevenly increasing the ballast settlement
in the transition zone also produced tie-ballast gaps and concentrated the wheel load on the
been extensively studied (Zaman et al. 1991, Briaud et al. 1997, White et al. 2005, and
Puppala et al. 2009), only limited research studies have focused on mitigating the
differential movement at railway track transitions. These few studies investigating track
presented test section and parametric analysis results on the effectiveness of these remedial
measures (Sasaoka et al. 2005, Read and Li 2006, and Nicks 2009).
differential movement problems at track transitions. Nicks (2009) divided the remedial
measures aimed at mitigating bump development at railway bridges into the following
30
interrelated categories: (1) reduce approach settlement, (2) decrease modulus on bridge
deck, (3) increase modulus on approach track, (4) reduce ballast wear and movement, and
(5) increase damping on the bridge deck. Kerr and Moroney (1993) concluded that most
problems at track transitions arise from rapid changes in the vertical acceleration of wheels
and cars in the transition zone. Accordingly, all remedial measures were recommended to
aim at reducing the train vertical acceleration at the transition zones, and could be divided
into the following three categories (Kerr and Moroney 1993): (a) smoothing the track
stiffness (often represented as ‘k’) distribution on the “soft” side of the transition; (b)
smoothing the transition by increasing the bending stiffness of the rail-tie structure on the
“soft” side, in close vicinity of the transition point; and (c) reducing the vertical stiffness
on the “stiff” side of the transition. Remedial measures under category (a) include use of
oversized ties, reduced tie spacing, ballast reinforcement using geogrids, hot-mix-asphalt
(HMA) underlayment, and use of approach slabs. On the other hand, the most commonly
known method under category (b) was developed by the German Federal Railways (DB)
and involves attaching four extra rails (two inside and two outside the running rails) to the
cross ties (Kerr and Bathurst 2001). Finally, the primary approach in category (c) involves
the installation of tie pads and/or ballast mats on ballast-deck bridges to reduce the track
stiffness on the “stiff” side of a transition point. Besides HMA underlayment, according to
Moale et al. (2016), Geoweb can be adopted as an alternative as it can reduce cost while
Chawla and Shahu (2016) found that the geogrid reinforcement under ballast layer
could significantly mitigate tie displacement only at low values of subgrade modulus and
effective subgrade shear strength parameters. It was equally effective at mitigating tie
31
displacement within the sub-ballast thickness from 450 to 1000 mm. The geotextiles at
subgrade surface could facilitate quick in-plane drainage and dissipation of pore water
Using analytical procedures, Kerr and Moroney (1993) engineered pad stiffness to
“match” the track running over the bridge to the approach track. Accordingly, a later study
(Kerr and Bathurst 2001) installed these “matched pads” on three different open-deck
bridges near Chester, PA, Catlett, VA and Philadelphia, PA in the US. Field test results
the bridge abutments. Sasaoka and Davis (2005) tried different methods to alter the track
stiffness and damping characteristics on bridge approaches. Installing ties made of different
materials, they reported that plastic ties on a concrete span ballasted-deck bridge effectively
reduced the stiffness difference at track transitions. Through parametric analyses using the
the approach and altering tie pad properties on the bridge deck were the most effective
dynamic analyses using NUCARS, they concluded that providing extra dampers on the
bridge deck could improve the impact attenuation at the transition by up to 30%. Li and
Davis (2005) concluded that remedies intended to strengthen the subgrade were not
effective for sites where ballast/subballast layers were primarily responsible for the
differential movement. In such cases, mitigation techniques such as rubber pads under the
concrete ties, or rubber mats on the concrete bridge deck had to be used to reduce the track
32
Rose and Anderson (2006) presented asphalt underlayment trackbeds as an
effective method for improving the performance of track transitions at tunnels, bridge
approaches, special trackwork like crossing diamonds, crossovers and turn-outs, as well as
the bridge and a thinner section close to the existing all-granular trackbed reportedly
improved performances of both open-deck and ballast-deck bridges. Rose and Anderson
(2006) reported four bridge approaches that were rehabilitated using this technique along
a Kentucky mainline with over 50 million gross tons (MGT) annual tonnage and a line
speed of 80-96 km/h. Over five years since the renewal of these approaches, no resurfacing
Concrete Fill (LCCF) with ground improvement to mitigate railroad bridge approach
settlement. Their project was conducted over San Diego River with a double track bridge,
the maximum settlement with normal weight fill is 4.8’’ while it’s only 1.9’’ with LCCF.
The maximum settlement with normal weight fill and compaction grouting was 1.4’’ while
it was only 0.5’’ with LCCF and compaction grouting. Therefore, lightweight fill is proven
useful for settlement mitigation although innovative materials can take significant time and
Apart from the above listed remedial measures, researchers have also suggested
approach slabs (Sharpe et al. 2002), and applying chemical grouting (Woodward et al. 2007,
Hyslip et al. 2009) and stone blowing (Chrismer 1990, McMichael and McNaughton 2003)
as alternatives to mitigate track transition problems. Bridge approach slabs have been used
33
as a semi-structural method to ease the transition from approach embankments to the fixed
bridge structure (Sharpe et al. 2002). However, concrete approach slabs require good
connection at the abutment and good support conditions away from the abutment to be
effective so they are not being widely used. Hyslip et al. (2009) proposed chemical grouting
Background and previous research study details related to chemical grouting, under
tie pads and stone blowing applications are listed next in Sections 2.6.1 to 2.6.3.
Chemical grouting, also known as polyurethane stabilization (see Figure 2.10), has
been successfully used in the past to reduce excessive ballast vibrations and long-term
monomers. The lower the closed cell content in polyurethanes, the more the material acts
like a flexible foam. In the case of a rigid-foam, where mechanical properties such as high
strength and stiffness are intended, for example, for ballast reinforcement, a high content
its negligible curing period and less interruption of traffic. Usually, no track shutdown is
required. Woodward et al. (2007) highlighted that when properly applied, chemical grouts
had the ability to bind the ballast particles together, thus reducing excessive vibrations and
does not adversely affect the drainage ability of the ballast layer.
34
Figure 2.10 Polyurethane Grouting of Ballast.
polyurethane coating of ballast, with the use of urethane only 1-2% by weight of dry ballast
aggregate, through a direct shear box test. Polyurethane coated ballast samples formed rigid
blocks and resulted in two to three fold increases in the shear strength while the amount of
ballast breakdown during testing decreased due to less particle movement and reorientation
under shear loading (Dersch et al. 2010 and Boler 2012). Boler (2012) further studied
ballast aggregates treated with polyurethane by the use of an image-aided particle shape
modeling and particle packing simulations using the Discrete Element Method (DEM) and
showed that shear strength increases of the urethane treated ballast assembly was
significantly influenced by the bond strengths of the particle contacts. Keene and Edil
(2012) tested specimens of polyurethane stabilized ballast under repeated loading. The
35
results showed that polyurethane stabilized ballast had good resistance to accumulation of
plastic strain compared to untreated substructure materials. Also, from the flexural strength
kPa) were close to that of cement stabilized soils. However, flexural modulus of
polyurethane stabilized ballast was much lower than that of cement stabilized soils. The
bonding and tensile capacity of polyurethane stabilized ballast are believed to provide
Besides laboratory testing, field testing and full-scale testing were also conducted
technique through laboratory tests under cyclic loading and then based on the good
performance from experiments, through a case study in Grovehill Tunnel, U.K. With
depths, gauge clearance issues on ballasted railway track could be maintained. Full scale
testing was conducted by Kennedy et al. (2013) to compare the traditional ballast and
cycles at different loading levels. Significant reduction in track settlement was achieved
with the polyurethane reinforcement. Polyurethane treated ballast railway track gave
an in-service railroad track in the State of Illinois, where track settlement and progressive
shear failure were found to be predominant on site. Though field settlements were not
measured, their study mainly focused on life-cycle cost analysis and the analyses related
to maintenance cycles of both the stabilized ballasted track and traditional track. Their
36
results showed that over a 10-year period, polyurethane injection could save them $1,000-
two bridge approaches along Amtrak’s NEC near Chester, Pennsylvania. According to
Figure 2.11, the peak transient deformation recorded in the ballast layer decreased
significantly after chemical grouting. It could be concluded that chemical grouting proved
support conditions under tie and reduce excessive vibrations and particle migration within
Figure 2.11 Effect of Chemical Grouting on Ballast Transient Deformation (Mishra et al.
2016).
37
2.6.2 Under tie pads
To maintain a smooth transitional stiffness along the track between regular track
and bridge deck and to slow down the track deterioration at the bridge approach, one of the
most commonly studied techniques is to use Under Tie Pads (UTPs), also known as Under
Crosstie Pads or Under Sleeper Pads (Lundqvist et al., 2006). UTPs are resilient pads
attached to the bottom surface of ties to provide an elastic layer between tie and the ballast.
They have been used in track structures for almost 25 years. According to Lundqvist et al.
using UTPs.
UTPs have been tested in laboratory by many researchers and practitioners (Le Pen
et al. 2018, Grabe et al. 2015). It is believed that with the use of UTPs, the contact area
between tie and ballast can be increased for concrete ties from around 3% to 10-35%
depending on the pad category. France and Austria utilizations of under tie pads proved
that tamping frequency can be lowered and therefore maintenance cost is greatly reduced
(Fimor 2015). Grabe et al. (2015) conducted full tie and half tie tests of UTPs under various
loading conditions such as static, dynamic, and cyclic loading and used tactile surface
sensors to measure the pressure under the tie. The results show that the use of UTPs can
around 10-33%, reduction of contact pressure by 70%, and reduction of ballast breakdown
(2005) conducted preliminary investigations on the benefits of using rubber pads under
concrete ties in a section of test track at the Transportation Technology Center (TTC) in
38
Pueblo, Colorado. The changing of track modulus and damping showed that UTPs not only
reduced track stiffness, but also increased track damping. UTPs could also reduce impacts
between ballast particles and concrete ties or bridge surfaces. Two different types of rubber
pads used as UTPs were tested by Sasaoka and Davis (2005). The pads were adhered to
the bottom of concrete ties and installed on both a concrete span bridge and a steel beam
span bridge. They concluded that both types of pads were successful at lowering the bridge
modulus to below that of the approach track as well as providing damping for the bridge
structure. Sasaoka and Davis (2005) suggested using durable material with long service
life as UTPs since it had to resist abrasive action of the ballast and exposure to elements
UTPs (Schneider et al. 2011). The UTP-installed track typically generated higher rail and
crosstie accelerations, but lower UTP strains. UTPs could also slow down geometry
ballast deck concrete tie bridge back in 2007 and subjected to heavy axle load traffic since
then. The bridge did not need any localized repair or major track maintenance since 2007
although maintenance efforts were required on a quarterly basis on this bridge before.
Moreover, high amplitude vibrations were significantly reduced. Pen et al. (2015) used
conclude the potential benefits of UTPs such as to: (1) increase the number and area of
contacts; (2) reduce the rate of plastic settlement; (3) reduce the support stiffness and spread
the load along a greater length of track; and (4) and add in a consistent increment to the
39
UTPs usually consist of two layers: one elastic spring layer and a geotextile to
protect the spring layer. To study the properties and effectiveness of under tie pads
numerical models have been developed. For example, a parametric study was conducted
rail pads, under tie pads, and ballast materials and the results were analyzed through
simulated responses using two numerical models. Schematic drawing of one of the
analytical models, known as GroundVib, is shown in Figure 2.12. Results showed that
under tie pad stiffness only influenced the lower part of the frequency spectrum (lower than
250 Hz). Higher under tie pad stiffness was found to lead to lower rail bending moment,
and lower vertical rail displacement and velocity. No clear relationship between under tie
pad stiffness and rail acceleration was found. However, higher under tie pad stiffness would
40
Witt (2008) examined the influence of UTPs using the FEM. Three types of UTPs
(stiff, medium, and soft) were used. Results showed that stiff UTPs did not influence
transition area load. The peak load caused by hanging crosstie could even increase when
using stiff UTPs as compared to using no pads. Similarly, when using soft UTPs, the force
on transition area did not exhibit large difference with no pad areas. The best results came
from using medium stiffness UTPs. Small variations of contact forces occurred and the
peak caused by the transition area disappeared. Insa et al. (2011) also performed
simulations to study the effectiveness of using UTPs. Obtained results showed that UTPs
bridge. Ribeiro et al. (2015) used FEM modeling to show that soft UTPs would amplify
rail displacement, tie accelerations, and reduce abrupt variations in track vertical stiffness.
Despite many of the successful installations of UTPs in different countries and field
applications, some researchers emphasize the importance of more accurate knowledge still
since the variations in stiffness and quality are crucial to track settlement (Witt 2008,
Ribeiro et al. 2015). Paixao et al. (2015) reported their findings of two transitions
monitored which had similar box culverts. UTPs were added in one of the transitions.
Comparison results indicated that the UTPs influenced dynamic behavior, increasing
vertical flexibility and amplifying both rail displacements and crosstie accelerations. The
installation of UTPs, in contrary to the original expectation, did not provide a smoothing
stiffness of transition after all. Hence, Paixao et al. (2015) indicated that UTPs had a
significant influence on track dynamic behavior that was yet to be carefully studied and
41
well understood to design properly with its resilient properties as well as for the geometry
requirements and other arrangements along the transition zones to achieve expectations.
potential remedial measures at bridge approaches. It consists of the addition of stone (or
ballast) to the surface of existing ballast such that any gap between tie bottom and ballast
layer is closed. Originally developed by British Railways, the process of stone blowing
involves the following steps (Selig and Waters, 1994): (1) The geometry of the existing
track is measured; (2) the precise track lift required at each crosstie to restore it to an
acceptable geometry is calculated; (3) the volume of stone that needs to be blown beneath
the crosstie to achieve such a lift is deduced from the known relationship between volume
of added stone and residual lift; and (4) the track is applied stone blowing. For stone
blowing, the target elevation during stone injection is always set “higher” than the desired
elevation so that enough space is provided to have the stones adequately distributed under
the tie. This leads to a smooth transition between the bridge approach and the bridge deck.
This is because the freshly injected stones underneath the ties usually undergo some
amount of settlement under train loading. The stone matrix tends to attain a “stable”
configuration through particle rearrangement and packing. It is therefore common for the
track profile to gradually “move down” with loading immediately after stone blowing until
a stable configuration is attained. The procedure can be illustrated as follows in Figure 2.13.
42
Figure 2.13 Stone Blowing Procedure (Selig and Waters, 1994).
According to Figure 2.14 from Mishra et al. (2016), the peak transient deformation
recorded in the ballast layer decreased significantly after stone blowing. The conclusion
was that stone blowing could be proved to be effective as a remedial measure for reducing
ballast transient deformation, and it could also effectively reduce excessive rail lift-off
43
Figure 2.14 Effect of Stone Blowing on Ballast Transient Deformation (Mishra et al.
2016).
2.7 Summary
Track geometry problems at railway transitions have been well identified. All
design standards including European Rail Research Institute (ERRI), UIC code 719,
Design Standards for Railway Structures and Commentary, and AREMA involve the
designs of transition zones (Muramoto 2013). Although railway track transitions include
tunnels, special track work, and highway/rail at-grade crossings, the most common
transition problems are often encountered at bridge approaches. Due to drastic differences
in substructure and loading conditions, the tracks on a bridge deck undergo significantly
small deformations under loading compared to the approach tracks. This sudden change in
track deformation behavior at the transition point results in extreme dynamic loading
conditions and ultimately leads to rapid deterioration of the track and bridge structural
components.
44
This chapter first provided an introduction to railway track systems with an
explanation of track transitions zones and special emphasis on bridge approaches. Previous
research studies that focused on track transition zones were reviewed to identify major root
causes of the rough train ride and other cause-effect relations. Field instrumentation efforts
for monitoring railway track transitions with the intention to identify main causes were
summarized and practical solutions adopted by researchers and practitioners were listed as
remedial measures. Existing analytical and numerical models established for railway track
modeling under either static or dynamic loading conditions were also reviewed for
effectiveness. In the end, mitigation solutions were presented, and three chosen remedial
measures identified in the current research effort, i.e., chemical (polyurethane) grouting,
45
CHAPTER 3: QUANTIFICATION OF RAILROAD BRIDGE
APPROACH PERFORMANCE
The first task in this PhD study involved identification of problematic track bridge
By acquiring data from instrumented bridge approaches, ground truth validation cases were
established for the dynamic track model developed in this study. This chapter describes
these ground truth field studies and also deals with statistical analyses conducted on the
collected field data to quantify track transient response and performance trends at the
interfaces. The NEC is primarily a high-speed passenger railroad with occasional freight
traffic. On the NEC, the high-speed Acela Express passenger train operates up to a
maximum speed of 241 km/h (150 mph). This segment of NEC near Chester consists a
total of four tracks. Two of the tracks are maintained for high-speed passenger trains,
typically operating at 177 km/h (110 mph). The current research focused on monitoring the
track response and performance data due to Acela passenger trains operating on the 2nd and
3rd main lines. Acela passenger train weights or wheel loads, were relatively consistent
46
over time periods and the train geometry data were known. Thus, ACELA train loading
related instrumentation data were used to validate the proposed track-train-bridge model.
Historical geometry data were obtained from Amtrak spanning a 60-month period
from January 2005 to January 2010. In June 2012, Ground Penetrating Radar (GPR)
scanning along the track was conducted to identify substructure features. The analyses of
track geometry data and GPR data led to three worst bridge approach locations as far as
recurring differential movement problems were concerned, including bridges over Upland,
Madison, and Caldwell Streets. These three locations with frequent maintenance needs
Figure 3.2 represents ground penetrating radar (GPR) as well as geometry data for
Upland St. location. Note that GPR and geometry data for Madison St. and Caldwell St.
locations can be found in Appendix A. From Figure 3.2, significant deviations in the
vertical profile are clearly identified, shown in red color. It is indicated that a “bump” exists
47
at a near bridge location about 15 ft. from the north abutment. At 60 ft. from north abutment
in the open track location, the vertical profile becomes smoother as regular track is situated
away from transition zones. Therefore, the instrumentation locations were chosen at 15 ft.
(near bridge) and 60 ft. (open track) for monitoring track performance. Note that the “bump”
shown in Figure 3.2 is more obvious at bridge “entrance” where the train moves from the
approach onto the bridge deck, rather than exiting the bridge.
Figure 3.2 (a) GPR Scans, (b) Mid-Chord Offset Data, (c) Space Curve at Upland Street
Bridge Approach.
of 2012. Multidepth Deflectometers (MDDs) were selected to install and monitor the
movement of individual track substructure layers. As stated earlier, the MDD technology
was first developed in South Africa in the early 1980s to measure individual layer
48
deformations in highway pavements (De Beer et al., 1989). MDDs typically consist of up
fixed anchor buried deep in the ground (Scullion et al. 1989). The use of MDDs to monitor
railway track performance has been extensively pursued in South Africa (Grabe and Shaw
2010, Priest et al. 2010, Vorster and Grabe 2013) as well as in the U.S. (Sussmann and
Figure 3.3 illustrates the location of each MDD module typically placed at layer
interfaces in the track substructure. Note that the LVDT #1 was installed at the bottom of
the crosstie. The rest of the LVDTs (#2 - #5) were installed at track substructure layer
interfaces. The anchor is placed sufficiently deep in the ground (10 ft. below the crosstie)
in each layer is that the anchor does not move with time under train loading. According to
the results from geometry data and GPR data, a total of six MDD strings were installed at
the Amtrak NEC bridge approaches, two at each selected bridge approach. MDDs were
installed at 15ft. and 60ft. from the North abutment at Upland Street location; 12ft. and
60ft. from South abutment at Madison Street location to monitor both near bridge and open
track locations, respectively. At Caldwell Street location, two MDDs were installed at 80
ft. from the south abutment on the left and right ends of the same time to monitor open
track performance.
49
Figure 3.3 Schematic of Track Substructure Profile with MDD Modules Placed at
Individual Layer Interfaces
Figure 3.4(a) shows a photo of the individual LVDT modules used in the MDD
system. The LVDTs were of 4-mm full range measurement and specially developed for
MDD usage. Figure 3.4(b) illustrates the independent anchoring system of MDDs. The
inner core of bottom-most LVDT (No.5) is mounted directly to the bottom anchor. The
core for LVDT module No. 4 is directly mounted on module No.5. The same pattern is
followed for the rest of LVDTs. Hence, except for the bottom-most MDD module, all other
MDD modules have movable anchors. The deflection measured by each LVDT represents
the deflection in that particular layer. Note that as LVDT module No.1 is attached to the
bottom of crosstie, LVDT 1 measures the total deflection of ballast and crosstie.
50
LVDT # 1
d1
Core # 1
LVDT # 2
d2
Core # 2
LVDT # 3
d3
Core # 3
LVDT # 4
d4 Core # 4
LVDT # 5
Core # 5
d5
(a) (b)
Figure 3.4 (a) Photo of LVDT Module and (b) Independent Anchoring MDD System.
The drilling process before installation of MDDs was carried out in small
increments of 75-100 mm. Compressed air and high capacity vacuum cleaner were used to
clean the drilled hole. One of the primary tasks during the drilling process was to identify
the locations of track substructure layer interfaces. This would facilitate the mounting of
individual LVDT modules at the layer interfaces to measure the deformations of individual
track substructure layers. Through visual inspection of soil samples in the field, layer
interfaces were established upon noticing significant differences in the material type
removed from the drilled hole. Figure 3.5 shows the track substructure layer types and
51
thickness encountered at the Upland Street instrumentation location (sketches for Madison
(a) (b)
Figure 3.5 Substructure Layer Profile for a) Upland Street 15 ft. from the North
Abutment b) Upland Street 60 ft. from the North Abutment
Besides the MDDs, strain gauges were also mounted on the rail at the instrumented
sites to measure vertical wheel loads applied during the passage of a train. Dual-element
350 Ohm shear gauges were welded on the rail at the neutral axis. After installation, a
calibration frame was used to correlate applied vertical load levels to the voltages induced
in the strain gauges by use of a Wheatstone bridge circuit. Figure 3.6 shows the photos of
the strain gauges during and after installation. Note that strain gauges were placed on two
rail sections, named as Type I strain gauge and Type II strain gauge. The Type I strain
gauge measured wheel load on rail on top of crosstie; while the Type II strain gauge
52
Figure 3.6 Photos of Strain Gauges Welded on Rail in the Field.
summarizing the main characteristics, often with visual methods and observations. In this
At the instrumented Amtrak NEC bridge approaches, the predominant train traffic
is Acela Express passenger train, operating at 177 km/h (110 mph). Acela Express trains
consist of two locomotives, one at each end of the train and six passenger cars in between.
Figure 3.7 shows the bogie, axle, and wheel spacing for the Acela Express trains. As shown
in the figure, wheel spacing for Acela locomotive is approximately 2.8 m; and wheel
53
Figure 3.7 ACELA Locomotive and Passenger Car Axle Spacings.
involved periodic data acquisition from the instrumented locations to record permanent
represent movements in that corresponding substructure layer. Figure 3.8 presents the
measured track substructure settlement trends from 2012 to 2015 at Upland Street
instrumentation location. Each marker on the figure represents one data collection. It is
noticed that the top layer (LVDT No.1) at near bridge location (15 ft. from abutment)
to that at open track location (60 ft. from abutment). Maintenance was more frequently
carried out at near bridge locations to maintain proper vertical profile. Layers 2 to 5 show
54
(a)
(b)
Figure 3.8 Layer Settlement Trends at Upland Street Location (a) Open Track (60 ft.
from abutment) (b) Near Bridge (15 ft. from abutment)
55
3.2.3 Transient responses under train loading
transient responses at instrumentation locations were also measured to quantify the elastic
bounce back behavior of the substructure layers under loading. The transient layer
deformations and wheel load data collected in the field are presented next. The Upland
Street location was chosen as an example to analyze the dynamic responses at railroad
bridge approaches. Detailed transient response data measured at Madison and Caldwell
Within 3 years from 2012 to 2015, seven data collection trips were organized at
Amtrak NEC bridge approach locations at the Upland Street site. Figure 3.9 shows the
transient vertical deformation time history recorded by LVDT modules under a passing
train at open track location (60 ft. from abutment). The recorded deformations
corresponded to the 8-car Acela Express passenger train, including one locomotive at each
end of the train and six regular passenger cars. A total of 32 peaks were observed from the
time history. These peaks corresponding to the passage of each wheel over the
instrumented tie are quite distinguishable. Note that the locomotives created higher
important to note that at the open track location, the transient responses collected at
different times are similar in magnitude. The maximum vertical transient deformation
56
Figure 3.9 Upland St. Open Track Transient Vertical Deformations with Train Passage
Figure 3.10 shows the transient load time history recorded by strain gauges under
a passing train at the open track location (60 ft. from abutment). Note that the strain gauge
circuit installed on rail crib location clearly registered 32 peaks, corresponding to the 32
wheels in an Acela Express train operating along the NEC with two locomotives (one each
at the front and rear ends of the train) and six passenger cars. As expected, the locomotives
register higher load levels. At the open track location, the transient loadings collected at
different times are similar in magnitude. The highest measured load level at the open track
57
Figure 3.10 Upland St. Open Track Transient Loads on Crib with Train Passage
Besides the open track location, field data were also collected at the near bridge
location (15 ft. from abutment) for performance measurement. Figure 3.11 shows the
vertical transient deformation time history from 2012 to 2015 the at Upland Street near
bridge location. Unlike the open track data, the near bridge data showed great variability
deformation continued to increase from August 2012 (2.5 mm) to July 1st 2014 (8 mm).
Next, the maximum vertical deformation suddenly decreased to 1.6 mm on July 22nd of
2014, then up to 4 mm in May 2015. This sudden decrease in deformation was due to
tamping maintenance activity conducted at the Upland Street location. In addition, a clear
trend of increasing dynamic response was observed with time at the near bridge location.
From visual inspection, it is noticeable that open track and near bridge locations have
58
Figure 3.11 Upland St. near Bridge Transient Vertical Deformations with Train Passage
Similarly, Figure 3.12 shows the transient load time history recorded by strain
gauges under a passing train at the near bridge location. There are 32 peaks, corresponding
to the 32 wheels in an Acela Express train. The transient loads collected at different times
at the near bridge location are similar in magnitude. The highest load level is approximately
150 kN, which is higher than that recorded at the open track location (same train passage).
This is due to the fact that strain gauge captures both static and dynamic loading effects.
Though static loading of the train contributes the same amount of magnitude to the open
track and near bridge locations, dynamic impact load coming from the wheel-rail impact,
irregularity, etc. is higher at the near bridge location. Considering the relatively stable
magnitude in load levels from 2012 to 2015, it is anticipated that the increasing trend of
transient deformation (Figure 3.11) is mainly associated with change in track substructure
properties (weaker support over time). This finding is confirmed by previously presented
measured settlement data, indicating that the substructure layers develop larger permanent
59
settlement over time at near bridge locations, creating a gap between layers which caused
Figure 3.12 Upland St. near Bridge -Transient Loads on Crib with Train Passage
data over time, as illustrated in Figures 3.9-3.12, statistical analyses were also conducted
for (i) determining the maximum/peak deformation and (ii) estimating the 95% confidence
interval (CI) of the field data collected. Again, Upland Street field data are presented as an
example herein. Statistical analyses of transient responses for Madison and Caldwell street
determined from each dataset collected and graphed in Figure 3.13. Each marker represents
one dataset collected in the field. The dashed blue line with the triangle markers illustrates
60
the peak transient deformation over time at the near bridge location. The solid orange line
with the round markers shows the peak transient deformation over time at the open track
location. The dotted lines show the 95% CI for the two locations, respectively.
From Figure 3.13 the peak deformation under locomotive passage at the open track
location is nearly constant from 2012 until 2015. On the contrary, the peak deformation
under locomotive passage at the near bridge location is changing over time, and thus the
deviation is larger, leading to a larger range for confidence interval. The statistics of the
field data, including mean, standard deviation and 95% CI, of the peak transient
deformation at both locations are summarized in Table 3.1. Based on the statistical analysis,
it is concluded that the vertical deformation behavior at the near bridge and open track
locations are statistically different because the mean of one location does not fall within
the 95% confidence interval of the other location. This implies the difference in dynamic
9
Nearbridge
Jul 1-14
8 Opentrack
7
Peak Displacement(mm)
Jun-13
6
5
Jan-15
4 Jul 22-14 May-15
Jan-13
Aug-12
3
2
Aug-12 May-15
1 Jan-13 Jun-13 Jul 22-14 Jan-15
Jul 1-14
0
4/1/12 10/18/12 5/6/13 11/22/13 6/10/14 12/27/14 7/15/15
Time
Figure 3.13 Upland St. Peak Vertical Deformation over Time (2012-2015)
61
Table 3.1 Statistics of Peak Vertical Deformation (2012-2015)
Besides the deformation, the loadings on track structure were also recorded and
analyzed statistically. Note that strain gauges were mounted on two rail sections to measure
two types of loading. One is located on rail where there is crosstie support underneath
(Type I strain gauge). The other measures the force on the crib of the rail, which equals to
the dynamic loading on top of rail (Type II strain gauge). The peak load corresponding to
locomotive wheel passage was selected from the dataset and graphed with time in Figure
3.14. The dashed blue line with the triangle markers illustrates the peak load measured by
Type I strain gauge over time at the near bridge location. Solid orange line with round
markers represents the peak load measured by Type I strain gauge at the open track location.
The dotted lines are the high and low limits of the 95% CI for the two locations,
respectively.
The loading time history does not reveal clear trends as in the case of the
deformation time history (see Figure 3.13). The maintenance activities occurred in July
2014 at the near bridge location is associated with a decrease in force under locomotive
passage. The statistics of the peak force measured by Type I strain gauge are summarized
62
in Table 3.2. The results indicate a statistically significant difference in the rail load on top
140
Aug-12 Jul 1-14
Jun-13
120 Jan-13 Jan-15
May-15
Peak Force on Tie (kN)
100
Jul 22-14
80
60
Aug-12 Jul 22-14 May-15
Jan-13 Jun-13 Jan-15
Jul 1-14
40
Nearbridge
20 Opentrack
0
4/1/12 10/18/12 5/6/13 11/22/13 6/10/14 12/27/14 7/15/15
Time
Figure 3.14 Upland St. Peak Force on Tie of Locomotives over Time (2012-2015)
Table 3.2 Statistics of Peak Vertical Load Measured on Top of Tie (2012-2015)
Similarly, the data collected by Type II strain gauges (mounted on crib of the rail)
were analyzed and graphed in Figure 3.15 which shows the peak load corresponding to the
63
locomotive wheel passage measured by Type II strain gauges. Again, the loading time
history does not reveal a clear trend as in the case of the deformation time history (see
Figure 3.13). The maintenance activity occurred in July 2014 at the near bridge location
did not cause any significant change in loading behavior. The statistics in Table 3.3 indicate
80
60
40 Nearbridge
20 Opentrack
0
4/1/12 10/18/12 5/6/13 11/22/13 6/10/14 12/27/14 7/15/15
Time
Figure 3.15 Upland St. Peak Force on Crib over Time (2012-2015)
Table 3.3 Statistics of Peak Vertical Load Measured on Rail Crib (2012-2015)
64
3.4 Track Substructure Modulus Backcalculation
Previous analyses showed that the open track and near bridge locations were
statistically different from each other as far as transient deformation and loading trends
were concerned. Moreover, the change in track substructure properties over time was
noticeable, especially at near bridge locations. To quantify the substructure properties and
obtained from the three instrumented NEC bridge approaches of Amtrak passenger lines
near Chester, Pennsylvania, USA were used to analyze properties of different layers as well
GEOTRACK (Chang et al. 1980), a multi-layer elastic software program, was used
as a tool for estimating track substructure layer moduli from the field data. It allows the
loads, properties of rails, ties, ballast and underlying layers and geometry of the track
structure (Li and Selig, 1998). With the measured displacements and axle loads at the
assumptions and features of GEOTRACK program are listed in Table 3.4. Figure 3.16
65
Table 3.4 Features and Assumptions of the GEOTRACK Program (Chang et al., 1980)
66
Figure 3.16 Track Elements in GEOTRACK Model (Reproduced from Chang et al. 1980)
The first step during the analysis of a given track structure using GEOTRACK
involves the selection of relevant input parameters for different track components. Some
of the input parameters such as rail, tie and fastener properties, etc., remain constant across
different sites, whereas other input parameters such as substructure layer properties, etc.,
change from one site to another. Table 3.5 lists typical parameters used in GEOTRACK
analyses. Note that K1 and K2 are coefficients for the calculation of stress-dependent soil
moduli. As the soil moduli is simplified to constant in the assumption, K1 and K2 are set
to zero.
measured by the strain gauge circuits were used as inputs in the GEOTRACK software.
Peak transient deformations recorded by the individual LVDTs corresponding to the time,
when the leading wheel of the trailing locomotive was directly on top of the instrumented
67
tie, were used as the substructure layer deformations. Position of the second wheel load
was determined based on the axle spacing (obtained from the locomotive manufacturers)
and tie spacing patterns at the instrumented bridge approaches. For instances when the
second wheel location was between two ties, standard force and moment balance methods
E (MPa) 2.07E+05
I (mm4) 3.90E+07
E (MPa) 2.07E+04
68
Table 3.5 (Continued)
K1 & K2 0
In the analyses, the two peak loads were assigned on three ties. When the first load
was assigned to the first tie, and the second load was 2,800 mm away from the first load
per the axle spacing based on axle spacing. The allocated load on the 5th tie and 6th tie can
thus be determined from the force and moment balance equations. An example of detailed
Assume that the two peak loads were recorded as 134 kN and 126 kN, respectively.
Crosstie spacing on site is 609.6 mm. These two peak loads need to be assigned on three
crossties. First load was assigned to the first tie (134 kN). According to force and moment
equilibrium equations, the reaction force caused by the second load can be determined.
69
Note that the axle spacing is 2,800 mm. Hence, the number of ties between the first load
2800 𝑚𝑚
= 4.59
609.6 𝑚𝑚
Therefore, the second load lies between 5th and 6th crosstie.
The distance from 5th tie for the second load is:
The distance from 6th tie for the second load is:
The allocated load on the 5th tie can be determined from the following equilibrium equation:
𝑀𝑜𝑚𝑒𝑛𝑡 (𝑀) = 0
248
𝐿𝑜𝑎𝑑 𝑜𝑛 5𝑡ℎ 𝑇𝑖𝑒 = 126𝑘𝑁 × = 51.26𝑘𝑁
609.6
After determining the loading condition, the moduli of individual track substructure
layers were then determined by trial and error. A set of seed moduli was first assigned to
the track substructure layers and GEOTRACK was used to calculate deflections at the
between the calculated and measured deflection values at each depth, adjustments were
70
made to the individual layer modulus values. For example, if the calculated deflections
were larger than those measured in the field, the modulus values assigned to the track
substructure layers had to be increased for the next iteration. This process was repeated
until the deflections measured in the field and those predicted using GEOTRACK matched
with a tolerance level of less than 5%. Some of the simplifying assumptions made during
such iterative estimation of track substructure layer moduli using GEOTRACK are:
1. The ballast and subballast layers at the Madison Street 12-ft. and Caldwell Street
west locations were combined into one layer for the analyses;
2. The two silty clay layers (layers 3 and 4) at the Madison Street open-track (60 ft.
from the south abutment) location were combined into one layer;
3. The sandy loam and sand layers (layers 3 and 4) at the Upland Street near-bridge
location (15 ft. from the north abutment) were combined into one layer.
system comprises contributions from (1) movement of the tie before it encounters the
underlying ballast layer, and (2) movement of the tie-ballast system as a unit. Accordingly,
taking the transient displacements recorded by the top-most LVDT as representative of the
ballast layer deformation can lead to serious over-estimation of movements within the
ballast layer if tie gap exists. This in turn leads to significant under-estimation of the
To more clearly indicate this, the first set of iterative substructure layer modulus
estimations were carried out without eliminating the tie gap contributions from the LVDT
No.1 measurements. The resulting backcalculated layer modulus values for the Madison,
71
Caldwell, and Upland Street locations are presented in Table 3.6. As seen from Table 3.6,
assuming no gap underneath the tie leads to the estimation of significantly low ballast layer
modulus values (mostly close to 20 MPa). These values are significantly lower than typical
ballast layer modulus values reported in the literature (Sussmann and Selig, 2000), and
therefore need to be revised based on elimination of tie-gap contributions from the LVDT-
1 transient deformations. This in turn specified the importance of considering tie gap in
modeling process rather than assuming a full support condition for track superstructure.
Table 3.6 Backcalculated Track Substructure Layer Moduli Values before Eliminating
Tie Gap
72
Table 3.6 (Continued)
To calculate the track substructure layer moduli for a more accurate representation
of the field load-deformation behavior of the ballast layer, it was necessary to subtract the
73
tie gap magnitudes from the transient deformations recorded by the top-most LVDT in an
MDD system. Tie gap amounts were estimated by Progressive Load Threshold method.
Detailed progress of determining gap amount can be found in Boler et al. (2018).
Estimated tie gap amounts were subtracted from the peak transient displacements
recorded by the top-most LVDTs in order to calculate the “Corrected Layer 1 Deformation”
values, which can be said to better represent the ballast layer deformation under train
loading. Such corrected layer 1 deformation values for the Madison, Caldwell, and Upland
Street bridge approaches are summarized in Table 3.7. It is seen that near bridge locations
usually exhibit higher tie gap compared with the open track locations, which explains the
In what follows, Table 3.8 presents the revised track substructure modulus values
estimated using GEOTRACK after the removal of the tie-gap contributions from LVDT-1
measurements. Figures 3.17 – 3.19 clearly indicate that the field measured data and
GEOTRACK predicted data match well using the final set of layer moduli.
74
(a)
(b)
Figure 3.17 Comparisons of Calculated Deformations with Field Measured Data (a)
Madison 12 ft. near bridge location (b) Madison 60 ft. open track location.
75
(a)
(b)
Figure 3.18 Comparisons of Calculated Deformations with Field Measured Data (a)
Caldwell West location (b) Caldwell East location.
76
(a)
(b)
Figure 3.19 Comparisons of Calculated Deformations with Field Measured Data (a)
Upland 15 ft. near bridge location (b) Upland 60 ft. open track location.
77
Table 3.7 Corrected Layer 1 Deformations – Considering Tie Gap
Field Measured
Corrected Layer
Layer 1 Tie Gap
Location 1 Deformation
Deformation Estimated (mm)
(mm)
(mm)
12 FT 0.83962 0.5454 0.29422
Aug 2012
60 FT 1.9038 1.1209 0.7829
12 FT 1.1875 0.9219 0.2656
Nov 2012
60 FT 0.74028 0.5873 0.15298
12 FT 1.5979 1.3731 0.2248
Jan 2013
60 FT 0.88728 0.6085 0.27878
12 FT 1.7133 1.4275 0.2858
June 2013
60 FT 0.96554 0.7637 0.20184
Field Measured
Corrected Layer
Layer 1 Tie Gap
Location 1 Deformation
Deformation Estimated (mm)
(mm)
(mm)
West 1.52 0.978 0.542
Nov 2012
East 0.976 0.7099 0.2661
West 1.5 0.9252 0.5748
Jan 2013
East 1.04 0.7166 0.3234
West 1.51 1.007 0.503
June 2013
East 0.794 0.618 0.176
78
Table 3.7 (Continued)
(c) Upland Street Location
Field Measured
Corrected Layer
Layer 1 Tie Gap
Location 1 Deformation
Deformation Estimated (mm)
(mm)
(mm)
15 FT 1.712 1.5286 0.1834
Aug 2012
60 FT 0.4537 0.3435 0.1102
15 FT 2.6261 1.9836 0.6425
Nov 2012
60 FT 0.34659 0.2364 0.11019
15 FT 1.594 1.3242 0.2698
Jan 2013
60 FT 0.40564 0.2819 0.12374
15 FT 4.5513 3.864 0.6873
June 2013
60 FT 0.31634 0.2176 0.09874
Table 3.8 Backcalculated Track Substructure Layer Moduli Values after Eliminating Tie
Gap
(a) Madison Street Location
Madison 12ft. (MPa) Madison 60ft. (MPa)
Layer Aug Nov Jan June Layer Aug Nov Jan June
2012 2012 2013 2013 2012 2012 2013 2013
Ballast 70 94 102 78 Ballast 32 155 96 105
Fouled 70 94 102 78 Fouled 56 80 80 60
Ballast Ballast +
Hard Pan
Hard Pan 37 51 57 40 Silty Clay 35 39 41 34
Grey 41 60 42 39 Silty Clay 35 39 41 34
Sandy
Loam +
Cinder
Brown 36 36 38 35 Fat Clay 54 50 54 49
Silty Clay
79
Table 3.8 (Continued)
(b) Caldwell Street Location
Caldwell West (MPa) Caldwell East (MPa)
Layer Nov Jan June Layer Nov Jan June
2012 2013 2013 2012 2013 2013
Ballast 52 57 50 Ballast 80 60 125
Fouled 52 57 50 Fouled Ballast + 60 49 50
Ballast Hard Pan
Comparing the estimated track substructure layer modulus values before and after
the consideration of the tie gap magnitudes indicates a significant increase in the estimated
ballast and subballast layer moduli for the latter case, which is more representative of
values reported in the literature. The estimated substructure layer moduli are particularly
80
useful as initial modulus estimates during advanced numerical modeling of railroad track
transitions.
3.5 Summary
results from advanced analyses of transient response data and vertical wheel load data
collected at the instrumented bridge approaches on the Amtrak Northeast Corridor near
Chester, PA. First of all, three bridge approach locations were selected for the
instrumentation effort based on analyses of GPR and 60-month period of geometry car data.
(MDDs) and strain gauges were installed at the selected bridge approaches to quantify
deformations of individual track substructure layers under loading as well as applied wheel
Next, exploratory analyses were conducted using the field data to visualize the time
history of dynamic responses in the track system. This step helps us to gain a better
that the vertical wheel loads at near bridge locations were relatively constant and stable
over the three years monitored, however, the deformations at near bridge locations showed
clear trends of increasing for the same monitoring period. This phenomenon indicates
noticeable changes in substructure support conditions over time at the near bridge locations.
In addition, statistical analyses were carried out on the field collected data. Mean,
standard deviation, and 95% confidence interval values were calculated for both the open
track and near bridge locations. It was found that the transient deformations and wheel
81
loads on rail from the two locations were statistically different. Such major differences
were not due to random error in data collection. There is a need to determine the physical
mechanisms that govern at railroad track bridge approaches and evaluate how they
Finally, a track substructure analysis tool GEOTRACK was utilized in this chapter
to quantify the individual moduli of substructure layers. An iterative process was developed
to backcalculate the layer moduli based on the field measured transient layer deformation
results and wheel loads. The unusual low moduli backcalculated based on LVDT-1 ballast
layer deformation typically indicated an insufficient support from underneath the tie or a
possible gap between the bottom of the tie and the top of the ballast layer. To better quantify
substructure layer properties, a gap forming at the tie-ballast interface was considered at
each instrumented site. After correction of the top most layer deformation due to this
estimated gap, the final backcalculated layer moduli were more reasonable in scale and
82
CHAPTER 4: DISCRETELY SUPPORTED TRAIN-TRACK-BRIDGE
MODEL
4.1 Introduction
Field collected transient deformation and wheel load data presented in Chapter 3
are used in this chapter to analyze the dynamic behavior of train-track-bridge coupling
interaction, which is a complex problem that accounts for when the train passes on the track
transition zone from an open track on to bridge deck. To gain a better understanding of the
dynamic behavior, and to quantify important variables according to their relative influences,
the numerical modeling technique has been adopted herein to develop an innovative model
to simulate the dynamic behavior in time domain at track bridge approaches. This chapter
will introduce this 2-D dynamic train-track-bridge model consisting of a coupled train and
track structure, which considers both regular track on substructure and bridge deck.
illustrated first. Next, the mathematical derivations of equilibrium equations are presented
for the train-track-bridge system and numerical solution algorithms are discussed. Finally,
nonlinear springs are introduced in the model to better represent the force-displacement
relationship at track bridge approaches with hanging crossties. Accordingly, the previously
adopted numerical solution algorithm is modified to solve the new nonlinear train-track-
bridge system.
model. The train is assumed to travel in the direction of arrow at a constant speed of 𝑣. The
train presented in the study consists of two locomotives at each end of the train (front/rear)
83
and six passenger cars in the middle, similar to the Amtrak’s ACELA Express passenger
Each vehicle consists of one car body, two bogies, and four wheel-sets. Each
component is assumed to be a nondeformable rigid body that only has vertical movement
of vehicle car body vertical displacement 𝑈F , vehicle car body pitch motion 𝜑F , two bogie
masses vertical displacement 𝑈HIJ , two bogie pitch motions 𝜑HIJ , and four wheel masses
vertical displacement 𝑈K . The vehicle body is linked with bogies through secondary
suspension springs (𝑘LM ) and dampers (𝑐LM ). The bogies are linked with wheels through
primary suspension springs ( 𝑘LN ) and dampers ( 𝑐LN ). The primary and secondary
Track substructure on the open track side is modeled as an infinite Euler beam
rail pads, ballast, and soil, and two layers of masses, representing crossties and ballast.
Crossties and ballast masses are assumed rigid bodies. The proposed model accounts for
shear stress between ballast particles. While track substructure on the bridge side is
foundations representing crossties resting directly on top of bridge deck (open deck bridge).
At each time step, the vertical displacement of rail 𝑈O , vertical displacement of crosstie 𝑈P ,
and vertical displacement of ballast masses 𝑈H are the remaining unknowns in time domain
84
85
Figure 4.1 Proposed Train-Track-Bridge Model.
To summarize, the proposed train-track-bridge model given in Figure 4.1 has the
following assumptions:
2) Each vehicle consists of one rigid car body, two rigid bogies, and four rigid wheel-sets;
3) The primary and secondary suspension systems in the vehicle are considered linear;
4) The vehicles are not connected (independent moving) in the train model;
5) The interaction between wheel and rail is assumed to be linked by linear springs in
7) All the springs and dampers in track substructure are assumed linear at the beginning;
To obtain realistic dynamic responses in the track system from the proposed train-
track-bridge model, the model parameters as input properties need to be chosen with
physical relevance and careful attention. Table 4.1 presents the physical meanings of the
notation symbols in the proposed train-track-bridge model. Table 4.1 also lists one set of
representative property values. Some of the track properties, such as crosstie spacing and
rail properties, are selected based on realistic values at instrumented bridge approach
locations at Amtrak NEC discussed in Chapter 3. The other track properties, such as spring
stiffness and damping ratio are all selected from a list of previously published studies which
also established a three-layer track model (Cantero et al., 2016). Vehicle parameters such
as mass of train, bogie and wheel spacings are selected from Amtrak Acela Express
passenger train’s vehicle data documented in research report published by Judge et al.
(2018).
86
Table 4.1 Notations and Values for Train-Track-Bridge Model in This Study
Track Properties
𝐸 Young's modulus of rail 2.07E+05 Mpa
𝐼 second moment of area 3.9E+07 mm4
mT rail mass per unit length 67.46 kg/m
𝑐U rail pad damping ratio 124 kN.s/m
𝑘U rail pad stiffness 7.80E+04 kN/m
𝑚P crosstie mass 386 kg
𝑚H ballast mass 683 kg
𝑐H ballast damping ratio 82 kN.s/m
𝑘H ballast stiffness 1.20E+05 kN/m
𝑡L ballast shear stiffness 7.80E+03 kN/m
𝑐L subgrade damping ratio 300 kN.s/m
𝑘L subgrade stiffness 5.00E+04 kN/m
𝑈O rail displacement -
𝑈P crosstie displacement -
𝑈H ballast displacement -
𝑘VO vehicle-rail stiffness 1.53E+06 kN/m
Vehicle Properties
𝑀F car body mass 6.31E+04 kg
𝐽F car body mass inertia 1.20E+06 kg.m2
𝜑F car body pitch rotation -
𝑈F car body displacement -
cYM secondary suspension damping 70 kN.s/m
𝑘LM secondary suspension stiffness 5.30E+03 kN/m
𝑀HIJZ[ bogie mass 1.20E+03 kg
𝐽HIJZ[ bogie mass inertia 760 Kg.m2
𝜑HIJZ[ bogie pitch rotation -
𝑈HIJZ[ bogie displacement -
𝑐LN primary suspension damping 49 kN.s/m
𝑘LN primary suspension stiffness 2.10E+03 kN/m
𝑚K wheel mass 1.20E+03 kg
𝑈K wheel displacement -
𝑃 wheel-rail interaction force -
𝑎 bogie distance 5.37 m
𝑏 wheel distance 1.42 m
87
4.3 Equations of Motion of the Train-Track-Bridge System
track, and bridge. To solve for the dynamic responses of these components, equilibrium
equations are derived for each component. Note that the individual vehicle is considered
independent. Hence, one vehicle rather than the whole train (eight vehicles) will be taken
motion of vehicle, track on embankment side, and track on bridge side are presented in an
orderly fashion and by discussing the assembly process of coupling these components.
Only vertical displacement and pitch motion are considered for the vehicle. The
body diagrams of dynamic force equilibrium for vehicle components are illustrated in
Figure 4.2. For example, the car body is supported by two bogies underneath. It reaches
dynamic equilibrium with the external forces from bogies and self-weight. Similarly, each
bogie is connected with the car body and two wheel-sets. Therefore, the force equilibrium
for the bogie is an equilibrium between self-weight of bogie, contact forces between bogie
and wheel-set 1, contact forces between bogie and wheel-set 2, and contact forces between
car body and bogie. Each wheel-set has self-weight and two external forces, including
contact force between wheel-set and rail and contact force between wheel-set and bogie.
All the forces exerted on wheel-set should reach equilibrium at any given time. Besides the
force equilibrium, the moment equilibrium for the bogie should also be achieved at the
same time.
88
Figure 4.2 Body Diagram for Vehicle Components.
The unknown dynamic responses in the model i.e. vertical displacement and pitch
rotation angles are in time domain. The system reaches both force and moment equilibrium
at any given time 𝑡 in the analysis period. Accordingly, the equations of motion for each
𝑀^ 𝑈̈^ + 2𝑐YM 𝑈̇^ + 2𝑘YM 𝑈^ − 𝑐YM b𝑈̇cdefgN + 𝑈̇cdefgM h − 𝑘YM b𝑈cdefgN + 𝑈cdefgM h
(4.1)
= 𝑀F 𝑔
𝐽^ 𝜓̈^ + 2𝑐YM 𝑎M 𝜓̇^ + 2𝑘YM 𝑎M 𝜓^ − 𝑐YM 𝑎b𝑈̇cdeN + 𝑈̇cdeM h − 𝑘YM 𝑎b𝑈cdeN − 𝑈cdeM h
(4.2)
=0
𝑀HIJZ[ 𝑈̈cdefgZ + (𝑐YM + 2𝑐YN )𝑈̇cdefgZ + (𝑘YM + 2𝑘YN )𝑈cdefgZ ∓ 𝑐YM b𝑈̇^ + 𝑎𝜓̇^ h
= 𝑀HIJZ[ 𝑔 (4.3)
89
𝐽^ 𝜓̈cdefgZ + 2𝑐YN 𝑏 M 𝜓̇cdefgZ + 2𝑘YN 𝑏 M 𝜓cdefgZ − 𝑐YN 𝑏b𝑈̇mN − 𝑈̇mM h
where 𝑃Z represents the contact force between wheel-set 𝑖 and rail. It is also a function of
time 𝑡 as the other unknowns in the model such as 𝑈KZ , 𝑈HIJZ[Z , etc. For simplicity, the
contact force between wheel-sets and rail are assumed to be simulated by a linear spring
connecting the two components. Hence, at any given time 𝑡 , 𝑃Z (𝑡) = 𝐾VO × Δ𝑑 =
𝐾VO × (𝑈O − 𝑈KZ ), where the magnitude of constant contact stiffness 𝐾VO is selected from
the Yu and Mao’s (2018) stochastic dynamic model. Their model established a 3D train-
track-bridge coupled system with refined wheel/rail interaction. The contact force 𝑃Z (𝑡) is
substituted into Equation 4.5. The new equilibrium equation of wheel-set 𝑖 is given by:
on top of ballast layer for the track embankment side. The rail is modeled as an infinite
Euler beam. The crossties are modeled as rigid masses at a spacing of 2𝑓𝑡. The ballast
layer is modeled as rigid masses at the same crosstie spacing (2𝑓𝑡). At track embankment
side (open track location), it is assumed that all the stiffnesses are linear. Therefore, the rail,
crossties, and ballast masses in the model are connected with linear springs and dampers.
90
Similar to the train component, the track structure components including rail, crossties, and
Figure 4.3 shows the free body diagrams of track structure components. The
crosstie comes to equilibrium under contact forces between rail and tie, contact forces
between tie and ballast, and self-weight. A portion of the rail is taken as an example to
illustrate the forces exerted on it, including vertical supports from underneath crosstie and
vertical wheel loads from one wheel-set running on top of that portion of the rail. Ballast
contact forces between ballast and subgrade, contact forces between ballast and crosstie,
and self-weight.
Assume that 𝑥 represents the location in special domain along the train moving
direction. Note that the vehicle is running at a constant speed 𝑣, therefore, the wheel-set
contact force 𝑃Z is moving along the rail with a constant speed 𝑣. Detailed equilibrium
91
y t
𝜕 t 𝑈O (𝑥, 𝑡) 𝜕 M 𝑈O (𝑥, 𝑡)
𝐸𝐼 + 𝜌O + v 𝑎w (𝑡)𝛿(𝑥 − 𝑥w ) = v 𝑃Z 𝛿(𝑥 − 𝑣𝑡 − 𝑥Z ) (4.7)
𝜕𝑥 t 𝜕𝑡 M
N ZzN
𝜕 M 𝑈P (𝑥w , 𝑡)
𝑚P = 𝑎w (𝑡) − 𝑏w (𝑡) (4.9)
𝜕𝑡 M
Equation of motion of the shear force between ballast mass is given by:
Similar to the open track embankment side, the rail is also modeled as an infinite
Euler beam for the bridge side (the Euler beam is extending from open track side to the
bridge side). In addition, since the field instrumented bridge is an open deck bridge (no
ballast), ballast layer is not considered on bridge deck in the proposed train-track-bridge
model. Therefore, on the bridge side in the model, the rail is supported by discrete crossties
directly sitting on top of a rigid bridge deck. The crossties are modeled as rigid masses at
92
a spacing of 2𝑓𝑡 between each crosstie. At first, since no hanging ties are considered, the
rail, crossties, and nondeformable bridge deck are assumed to be connected by linear
for bridge abutment in the field. Realistically, if the bridge deck spans long distance, it
would undergo large deformation under dynamic loading. However, the proposed model
focuses on the dynamic responses at open track location and near bridge location.
Therefore, the bridge deck behavior is simplified. The track structure components
including rail and crossties reach the state of force equilibrium at any given time 𝑡.
𝜕 M 𝑈P 𝜕𝑈P (𝑥w , 𝑡)
𝑚P M
= 𝑎w (𝑡) − 𝑘𝑈P (𝑥w , 𝑡) − 𝑐 (4.13)
𝜕𝑡 𝜕𝑡
The above equilibrium equations 4.1 – 4.13 for the vehicle, track, and bridge system
can be solved numerically with reasonable time steps and initial conditions. In solving
these equations, the partial differential equations of motion of the rail (Equation 4.7) need
to be facilitated first using a reasonable shape function. This process is explained in detail
93
4.3.4 Weak formulation of rail beam equation
To solve the 4th order partial differential equation of motion of the rail beam, the
weak formulation is developed with utilization of rail mode shape function. The admissible
𝑘𝜋𝑥
𝜑€ (𝑥) = 𝑠𝑖𝑛 ( ) (4.14)
𝑙
where
Next, weak formulations are performed on the rail beam equation of motion to
obtain a series of 2nd order ordinary differential equations. Equation 4.7 is transformed as:
‡ y
𝜕 t 𝑈O (𝑥, 𝑡) 𝜕 M 𝑈O (𝑥, 𝑡)
… †𝐸𝐼 + 𝜌O + v 𝑎w (𝑡)𝛿(𝑥 − 𝑥w )
𝜕𝑥 t 𝜕𝑡 M
ˆ N
− v 𝑃𝛿(𝑥 − 𝑣𝑡 − 𝑥‰ )Š 𝑤Z 𝑑𝑥 = 0
‰zN (4.16)
where
𝑤Z = 𝜑Z (𝑥) (4.17)
described in the form of matrix. The matrix describing equations of motion of rail beam
where
‡ (4.19)
[𝑀N ]Z‰ = … 𝜌O 𝜑Z (𝑥)𝜑‰ (𝑥)𝑑𝑥
ˆ
‡ y
The equations of motion for vehicle components, crossties and ballast masses are
all used to assemble the final matrix. To be concise, one car of the train, consisting of one
car body, two bogies, and four wheel-sets were considered here to illustrate the basic form
[𝑀M ]“𝑈̈P ” + [𝐶• ]{𝑞̇ } + [𝐾• ]{𝑞 } + [𝐶t ]{𝑈Ṗ } + [𝐾t ]{𝑈P } + [𝐶– ]{𝑈Ḣ } + [𝐾– ]{𝑈H }
= {0} (4.26)
[𝑀• ]“𝑈̈H ” + [𝐶— ]“𝑈̇H ” + [𝐾H ]{𝑈H } + [𝐶˜ ]{𝑈Ṗ } + [𝐾˜ ]{𝑈P } = {0} (4.27)
where
95
š
[𝑈V ] = “𝑢F , 𝑢HIJZ[N , 𝑢HIJZ[M , 𝑢KN , 𝑢KM , 𝑢K• , 𝑢Kt , 𝜑F , 𝜑HIJZ[N , 𝜑HIJZ[M ” (4.28)
[𝑀V ] = 𝑑𝑖𝑎𝑔(𝑚F , 𝑚HIJZ[N , 𝑚HIJZ[M , 𝑚KN , 𝑚KM , 𝑚K• , 𝑚Kt , 𝐽F , 𝐽HIJZ[N , 𝐽HIJZ[M ) (4.29)
96
𝑡L −𝑡L 0 … 0 0
⎡−𝑡 2𝑡 −𝑡L 0 ⎤
L L … 0
⎢ ⎥
0 −𝑡 0 0 ⎥
[𝐾— ] = 𝑑𝑖𝑎𝑔(𝑘LN + 𝑘HN , … , 𝑘Lw + 𝑘Hw ) + ⎢ L … (4.42)
⎢ 0 0 −𝑡L 0 ⎥
⎢ 0 … 0 −𝑡L 2𝑡L −𝑡L ⎥
⎣ 0 0 … 0 −𝑡L 𝑡L ⎦
In solving all the above equations, the general equations of motion in standard
where
{𝑋} denotes the generalized displacement vector containing displacement of car body,
bogies, wheel-sets, rail, crossties, and ballast masses, as well as the pitch angle of car body
𝑀V 0 0 0
0 𝑀N 0 0
[𝑀] = § ¨
0 0 𝑀M 0
0 0 0 𝑀•
𝐶V 0 0 0
0 𝐶N 𝐶M 0
[𝐶 ] = § ¨
0 𝐶• 𝐶t 𝐶–
0 0 𝐶˜ 𝐶—
𝐾V ′ 𝐾VO 0 0
[𝐾] = §𝐾VO 𝐾N ′ 𝐾M 0
¨
0 𝐾• 𝐾t 𝐾–
0 0 𝐾˜ 𝐾—
97
{𝐹 } denotes the generalized force vector;
{𝐹 } = {𝐹V‘ , 0,0,0}š
Note that 𝐾V‘ , 𝐾N‘ , 𝐾VO , 𝐹V ′ are modified matrix according to the vehicle-rail contact force 𝑃Z .
presented in section 4.3. It is important to note that the generalized stiffness matrix needs
to be updated in each iteration to simulate the effect of moving vehicle (location of vehicle-
There are several widely used numerical integration methods for solving the
generalized matrix. For instance, Fourier transformation is one method to solve linear
leads to harmonic components 𝑃(𝑖𝑤). The frequency domain harmonic responses are
simply the product of the harmonic components by the receptance Frequency Response
Function (FRF) 𝐻(𝑖𝑤). The total response is the summation of the harmonic responses by
The Fourier transformation has been adopted here to solve a simplified model,
simulating one wheel on the track substructure moving from open track location to bridge
deck, as a base for evaluating numerical integration effectiveness. The detailed derivation
and numerical results for Fourier transformation integration can be found in Appendix C.
stiffness properties vary in time. In this situation, the calculation of the response of
98
nonlinear system can be done only by step by step direct integration. In the track-bed
system, the suspension system between crossties and ballast particles can be nonlinear if
the crossties are not properly supported: no contact force at the beginning of loading until
In Chapter 3, it was stated that field monitoring data revealed unsupported crossties
loading. Many methods exist for the direct integration of the equations of motion. Herein,
Newmark’s method has been chosen to solve the equations of motion in the track bed
Newmark (1959) for use in structural dynamics, based on finite difference method.
In this section, the algorithm for solving linear track system shown in Equation 4.43
is presented.
Table 4.2 illustrates the initialization and integration steps in the algorithm. Average
unconditionally stable and is always converging in the end. Therefore, no time step
restriction applies. The time step of ∆𝑡 = 0.001 can be selected in this case to account for
a compromised accurate loading function and computational time. The initial conditions
99
Next, the algorithm is implemented in MATLAB for numerical solutions. The
Table 4.2 Algorithm for Numerical Integration using Newmark’s Method at Linear
System (Paultre, 2013)
𝜸 𝜸
𝒂𝟒 = − 𝟏; 𝒂𝟓 = { − 𝟏| ∆𝒕; 𝒂𝟔 = (𝟏 − 𝜸)∆𝒕; 𝒂𝟕 = 𝜸∆𝒕
𝜷 𝟐𝜷
¼ = 𝒌 + 𝒂𝟎 𝒎 + 𝒂𝟏 𝒄
2. Calculate effective stiffness: 𝒌
Integrate step by step
for 𝒏 = 𝟎, 𝟏, 𝟐, … , 𝒕𝒅 /∆𝒕
100
4.5 Consideration of Hanging Crossties at Bridge Approach
As discussed in the previous section, the ballast layer stiffness (𝑘Hć‡ÄLP ) and
damping ratio (𝑐Hć‡ÄLP ) are assumed to be linear at open track locations. To clarify any
misunderstanding, in this Chapter, “open track location” refers to a location far from bridge
abutment and is assumed not to be affected by the discontinuity in transition zone; while
“near bridge location” refers to a location that is close to bridge abutment and may develop
transition problems. The linear track system at open track locations can be solved by the
However, very often at near bridge locations, hanging crossties exist in the track.
These hanging crossties commonly lack adequate support from the underneath ballast layer.
In Chapter 3, Table 3.7 lists the typical magnitudes of gaps observed between instrumented
crossties and top of ballast layer in Amtrak’s Northeast Corridor lines near Chester, PA.
to the gap between tie and ballast. For simplicity, the proposed nonlinear train-track-bridge
that at the beginning of loading, there is no force between crosstie and ballast until the gap
(𝑢J ) is closed. Once the distance between crosstie and ballast reaches the magnitude of gap,
the stiffness of springs between crosstie and ballast becomes linear (𝑘′Hć‡ÄLP ). Note that
the tangent of stiffness from open track and near bridge locations might be different:
constant in the analysis. Figure 4.4 illustrates the force-displacement relationship at (a)
open track location with linear stiffness and (b) near bridge location with nonlinear stiffness.
101
The gap between crosstie and ballast is represented by 𝑢J . Note that in the model, it is
flexible to choose the locations of the gap since it is discretely supported. The amount of
gap underneath each crosstie can also be assigned arbitrarily based on realistic situation.
(a) (b)
Figure 4.4 Linear and Nonlinear Dynamic Systems: (a) Linear Stiffness at Open Track
Location, (b) Nonlinear Stiffness at Near Bridge Location.
Based on the Newmark’s method for the linear system presented in Section 4.4, a
modified numerical algorithm is proposed for the nonlinear system. The track system at
near bridge locations may develop nonlinear stiffness that depends on the displacements in
time. Reduction of the time step ∆𝑡 leads to reduction in the error within a time step. Note
that in solving the nonlinear system, the tangent stiffness needs to be evaluated at each
iteration, which would considerably increase the calculation time. Before the gap between
102
tie and ballast closes, ballast layer stiffness remains small (avoid zero for computational
purpose). Once the gap closes, ballast layer stiffness becomes constant. Therefore, an
presented in Table 4.3 and utilized to solve the nonlinear track system proposed for near
bridge location.
Table 4.3 Algorithm for Numerical Integration using Newmark’s Method for Nonlinear
System (Bilinear)
𝟏 𝜸 𝟏 𝟏
𝒂𝟎 =; 𝒂 𝟏 = ; 𝒂 𝟐 = ; 𝒂 𝟑 = − 𝟏;
𝜷∆𝒕𝟐 𝜷∆𝒕 𝜷∆𝒕 𝟐𝜷
𝜸 𝜸
𝒂𝟒 = − 𝟏; 𝒂𝟓 = { − 𝟏| ∆𝒕; 𝒂𝟔 = (𝟏 − 𝜸)∆𝒕; 𝒂𝟕 = 𝜸∆𝒕
𝜷 𝟐𝜷
à at time 𝒕𝒏}𝟏 :
7. Calculate effective force matrix 𝒑
à 𝒏}𝟏 = 𝒑𝒏}𝟏 + (𝒂𝟎 𝒖𝒏 + 𝒂𝟐 𝒖𝒏̇ + 𝒂𝟑 𝒖𝒏̈ )𝒎 + (𝒂𝟏 𝒖𝒏 + 𝒂𝟒 𝒖𝒏̇ + 𝒂𝟓 𝒖𝒏̈ )𝒄
𝒑
8. Calculate displacement matrix at time 𝒕𝒏}𝟏 :
𝒖𝒏}𝟏 = 𝒑 Ã 𝒏}𝟏 /𝒌¼
9. Calculate acceleration and velocity matrix at time 𝒕𝒏}𝟏 :
𝒖̈ 𝒏}𝟏 = 𝒂𝟎 (𝒖𝒏}𝟏 − 𝒖𝒏 ) − 𝒂𝟐 𝒖𝒏̇ − 𝒂𝟑 𝒖𝒏̈ ; 𝒖̇ 𝒏}𝟏 = 𝒖𝒏̇ + 𝒂𝟔 𝒖̈ 𝒏 + 𝒂𝟕 𝒖̈ 𝒏}𝟏
103
4.6 Summary
train-track-bridge model that simulates the vertical dynamic responses when a train vehicle
moves along the rail track structure at bridge approaches. At first, the various components
of the train-track-bridge model was illustrated and explained in detail. Assumptions in the
proposed model were summarized. Next the equations of motion of all the components of
the model were derived assuming that the system reaches dynamic equilibrium at every
time step 𝑡. These equations of motion were then combined to formulate a generalized
matrix.
Newmark’s numerical integration method was chosen to solve the set of ordinary
differential equations for the track system because it can be applied to both a linear system
(open track location) as well as a nonlinear system (near bridge location) with a tie-ballast
gap. The force-displacement relationship between crossties and ballast masses at both open
track and near bridge locations were presented. The spring stiffness was assumed to be
evaluate the tangent stiffness at every step to solve for the nonlinear system within tolerable
error.
vehicle, rail, crosstie, and ballast masses at different locations along the track, etc. Once all
the parameters (stiffness and damping coefficients) in the model are calibrated with field
data, this model can be utilized to predict and analyze the in-track behavior, especially with
104
proper consideration of nonlinear properties at near bridge locations, and to quantify and
105
CHAPTER 5: FIELD VALIDATION OF TRACK MODEL
approach behavior under complex dynamic train loads, and the algorithms for solving the
proposed numerical model were presented. In this chapter, the developed model is
calibrated and validated with field instrumentation data collected from both open track and
near bridge locations. The field instrumentation data are chosen carefully to select track
model input parameters. The validation process for open track location is conducted with
linear model parameters. Then, the near bridge location is studied with consideration of
a simplified bi-linear relationship depending on the field measured gap size between the
demonstrated to effectively simulate and analyze the dynamic responses of track structure
under moving train load. The model is presented as an analytical tool to provide insight
into optimizing track structure designs for improved track performance at bridge
approaches.
deformations and wheel loads were available at different times and locations. Data
interpretations for track properties were conducted and general trends were clearly
demonstrated in Chapters 3 and 4. This chapter intends to utilize the data collected at the
Upland Street location to calibrate the input parameters in the proposed track model.
106
Therefore, the field data hereafter in this chapter refers to data collected at Upland Street
In the field near Chester, PA, the ACELA Express passenger train operates at a
speed of 110 mph when passing the Upland Street location. It consists of eight cars in total,
two locomotives at each end of the train and six passenger cars in the middle. The passenger
car weight is approximately 139 kips. Figure 5.1 shows the ACELA Express axle
configuration. The dimensions are in meters. Input parameters for vehicle part in the
proposed model are selected according to the above mentioned field loading patterns
Transient deformation data with train passage were collected in the field at two
locations: open track location (60 ft. from bridge abutment) and near bridge location (15 ft.
from bridge abutment). Figure 5.2 shows the field deformation data collected during train
passage in January 2015. Note that the January 2015 data were taken as an example to
illustrate the large displacements at near bridge and contrast the trends between open track
and near bridge locations. Such large movements at the bridge approach would lead to
107
accelerated track geometry degradation. Instead of only choosing one set of data (e.g.
January 2015) for the validation purpose, the confidence intervals of field data calculated
from different times were also used to calibrate the input parameters in the proposed track
model. In Chapter 3, the 95% confidence intervals (CI) were calculated and documented.
At the open track location, the peak displacement caused by locomotive is between 1.55
mm and 1.71 mm (95% CI); the peak displacement caused by passenger car is between
1.14 mm and 1.28 mm (95% CI). At the near bridge location, the peak displacement by
locomotive is between 2.10 mm and 5.98 mm (95% CI); the peak displacement by
Figure 5.2 Transient Deformations Measured in the Field at Upland Street Location –
January 2015.
108
5.2 Model Validation for Open Track Location
To eliminate the boundary effects of vehicle entering and leaving the track structure,
the number of crossties on the embankment side is assumed to be 500 and the number of
crossties on the bridge side is also assumed to be 500. Crossties are located at a spacing of
0.6096 m (2 ft.). In the field open track location, the data were collected at 60 ft. from the
bridge abutment, approximately 30 crossties away from the bridge deck. Figure 5.3
The locomotive in this case consisted of one car body of 45,000 kg, two bogies of
1,200 kg and four wheels of 1,200 kg, representing a self-weight of 90,000 kg (198 kips)
car. The passenger car consists of one car body of 31,550 kg, two bogies of 1,200 kg, and
four wheels of 1,200 kg, representing a self-weight of 63,100 kg (139 kips) car. The vehicle
is moving at a speed of 50 m/s. The rail unit weight is 67.46 kg/m. The flexural rigidity of
rail beam is 8.073 𝑀𝑁 ∙ 𝑚M . The concrete crosstie mass is 386 kg. Equivalent ballast mass
is 683 kg. The stiffness and damping ratio values are determined by trial and error within
a reasonable range: Rail pad damping coefficient of 124 𝑘𝑁 ∙ 𝑠/𝑚M ; Rail pad stiffness of
Ballast shear stiffness of 7.8 𝑀𝑁/𝑚M ; Subgrade damping coefficient of 300 𝑘𝑁 ∙ 𝑠/𝑚M ;
109
Figure 5.3 Sketch of Response Observation Point in the Model – Open Track.
With the ACELA Express passenger train moving from embankment side to bridge
side in the model, the structural responses including transient deformation, reaction force,
velocity and acceleration can be obtained from the model. Transient deformations obtained
from the numerical model are used to validate with field collected data. Figure 5.4 shows
the comparisons between model results and field data. Figure 5.4 (a) clearly shows that the
results from the proposed model match well with the field data collected in January 2015.
The model simulation results demonstrate 32 peaks, which correspond to 8 cars with one
passage of ACELA passenger train. The front and rear cars register higher displacements
due to heavier weight of the locomotives. Figure 5.4 (b) compares the confidence intervals
of field data and model results. It can be seen that the maximum peak displacement value
from the model falls into the 95% confidence interval calculated from the maximum peak
displacement value from the field data collected between August 2012 and May 2015.
110
(a)
(b)
Figure 5.4 Comparisons of Model Results and Field Data – Open Track.
111
5.3 Model Prediction for Open Track Location
The comparisons between field data and model results imply that responses at the
open track location can be accurately reflected and simulated by the proposed vehicle-
Figure 5.5 shows the transient deformations at the open track location at different
substructure depths, including top of rail, top of crosstie and ballast layer. In this simplified
three-layer linear track structure system, ballast layer contributes approximately half of the
Figure 5.5 Transient Deformations in Different Substructure Layers with Train Passage –
Open Track.
112
The graphs shown in Figure 5.6 present the (a) rail-crosstie reaction force and (b)
crosstie-ballast reaction force with train passage at the open track location. Note that the
rail-crosstie reaction and the crosstie-ballast reaction forces are similar in both trend and
magnitude. This confirms that the force is correctly simulated in the model since at the
open track location the crossties are in good contact with the underlying ballast particles
and thus the external force exerted on crossties can be passed onto the ballast mass
adequately. Chapter 3 lists that the average measured wheel load on rail crib in the field at
the open track location is around 132 kN for locomotive and 91 kN for passenger car. In
model simulations, the crosstie reaction force is approximately 40 kN for locomotive and
29 kN for passenger car. Hence, it is found that each crosstie takes up to 30%-32% of the
113
(a) Rail-Tie Reaction Force
114
Besides the reaction force, the structural vibration velocities and accelerations can
also be obtained from the model simulation. Figure 5.7 shows the time history of vibration
velocities of ballast mass, crosstie and rail, respectively. The peaks represent each wheel-
set of the ACELA Express passenger train. Note that the vibration velocities of different
components in the system have similar magnitudes in the order of 10~M m/s. Furthermore,
ballast mass registers the lowest vibration velocity in the system of 0.027 m/s; crosstie
mass shows slightly higher velocity of 0.039 m/s; and rail mass implies the highest velocity
of 0.058 m/s. The time history of vibration accelerations of the track system are graphed in
Figure 5.8. The vibration accelerations of different components are all below one standard
gravitational constant 𝑔 = 9.8 𝑚/𝑠 M . The peak accelerations at the open track location are
approximately 2 𝑚/𝑠 M for ballast mass; 3 𝑚/𝑠 M for crosstie; and 4.5 𝑚/𝑠 M for rail mass.
Figure 5.7 Structural Vibration Velocities with Train Passage – Open Track.
115
Figure 5.8 Structural Vibration Accelerations with Train Passage – Open Track.
After successfully calibrating and matching the model results with the open track
location field data, further validation is conducted for the near bridge location. The model
input parameters are kept the same as those for the open track location except for the
observation point in the model and the hanging crosstie gap induced nonlinear stiffness
assignment.
At the field instrumented near bridge location, data were collected at 15 ft. from the
bridge abutment, approximately 7 to 8 crossties away from the bridge deck. Therefore,
crosstie No.493 is chosen as the observation point in the model as shown in Figure 5.9.
116
The time history of structural responses at the observation point are calculated to simulate
dynamic response behavior of the track structure at the near bridge location.
Figure 5.9 Sketch of Response Observation Point in the Model – Near Bridge.
representing hanging crossties should also be properly simulated. Hanging crossties were
observed in the field at the near bridge location, as explained in Chapter 3, where there
were gaps between these hanging crossties and the underlying ballast particles. At the
instrumented Upland Street location, the gap was found to increase over time when the
track lacked proper maintenance. For calibration and validation purposes, a typical gap of
1.5 mm developed in January 2013 is selected in the model simulation. Such gaps are
created underneath crossties No. 485 to No.500 in the model. Before the gap closes, it is
assumed that no force exists between crosstie and ballast mass; as soon as the gap fully
closes during the loading process, the ballast stiffness abruptly increases to the same as in
117
open track locations – 120 𝑀𝑁/𝑚M . Please note that for simplicity, it is assumed that the
ballast damping ratio remains constant at the near bridge location. Using the nonlinear
algorithm presented in Chapter 4, numerical solutions from the model can be obtained.
simulations are compared to the field measured transient deformations. Figure 5.10 shows
the comparisons between the model predictions and the field data. The model simulation
results clearly demonstrate 32 peaks, which correspond to 8 cars with one passage of
ACELA passenger train. The front and rear cars register slightly higher displacements due
to heavier self-weights of the locomotives. Due to the existence of gap, the peak
displacement of a moving locomotive can be as high as 2.6 mm. If the gap continues to
increase, it is anticipated that the transient deformation can reach to a significantly high
Figure 5.10 (a) shows the general trends of transient deformations in terms of peak
location and peak magnitude under loading that are comparable with the field data.
However, the deformations predicted between wheel axles do not match very well with the
field measured data. The physical meaning of this “delayed bouncing” is that in the
simplified physical relationship between displacement and force for hanging crossties,
there is a little reaction force observed when the differential movement between crosstie
and ballast mass is small (𝑘~0). Hence, there is not sufficient upward direction force
exerted on crosstie for the bounce back movement. While in the field, even though there is
no upward force from the underlying ballast, the fastening system will lead the crosstie to
118
Figure 5.10 (b) compares the confidence intervals of field data and model results.
The upper and lower boundaries of 95% confidence intervals are graphed for locomotive
passage and passenger car passage, respectively. Due to the increase in the gap amount
with years, the peak displacement may vary significantly. Therefore, the higher and lower
boundaries of the confidence intervals are relatively far away from each other compared
with the open track case. It is noticed that maximum peak displacement value from the
model still falls into the 95% confidence interval calculated from the maximum peak
displacement value from the field data collected between August 2012 and May 2015.
The proposed model is considered to have the predicton ability for the general
trends and peak responses although it was somewhat unable to predict the bouncing
movement. Considering that the main purpose of the model is to detect the critical
responses, it is believed that the developed model has been successfully validated with field
119
(a)
(b)
Figure 5.10 Comparisons of Model Results and Field Data – Near Bridge.
120
5.5 Model Prediction for Near Bridge Location
Figure 5.11 shows the transient deformations at the near bridge location at different
substructure depths, including top of rail, top of crosstie and ballast layer. Recall that with
the exact same input parameters at the open track location, the ballast layer contributes
approximately half of the total vertical deformation. While at the near bridge location, the
ballast layer only contributes 1/3 of the total vertical deformation. Ballast mass itself
deformations under the same loading as a result of inadequate support from underlying
ballast layer. These results further explain that in the three-layered track system consisting
of rail, crosstie and ballast, hanging crossties control the total vertical deformation under
train passage. It is of vital importance to mitigate these tie-ballast gaps that occur at the
121
Figure 5.11 Transient Deformations within Different Substructure Layers due to Train
Passage – Near Bridge.
Figure 5.12 shows the predicted rail-crosstie reaction force and the crosstie-ballast
reaction force obtained from the vehicle-track-model. Clear loading patterns can be
observed in the time history. However, the nonlinear model at the near bridge location
registers more “noise” when compared to the open track location. From the field measured
data, the average peak locomotive wheel load on rail crib is approximately 148 kN while
the average peak wheel load on rail crib due to passenger car is approximately 100 kN.
From the numerical model predictions, the peak rail-crosstie reaction forces corresponding
to locomotive and passenger car are found to be around 100 kN and 80 kN, respectively.
Even though the impact wheel load on crib is only slightly higher than that at the open
track location, the reaction force on crossties takes up to 67%-80% of the total impact force
at the near bridge location compared with 30% of the total force at the open track location.
122
Such high impact forces would lead to accelerated degradation of both rail and crossties in
the field.
In addition to the rail-tie reaction force, tie-ballast reaction force also exhibits quite
different behavior as compared to the open track location. At the open track location, the
rail-tie reaction and tie-ballast reaction forces are comparable in both trends and magnitude.
However, at the near bridge location, it is noticed that the tie-ballast reaction force is
considerably smaller than the rail-tie reaction force. The difference between the rail-tie
reaction force and tie-ballast reaction force is the total external force on crosstie, which
123
(a) Rail-Tie Reaction Force
124
Figure 5.13 shows the time history of vibration velocities of ballast mass, crosstie
and rail, respectively. Figure 5.14 illustrates the time history of vibration accelerations of
these components. The peaks in the time history are corresponding to the passage of
wheelsets of ACELA Express passenger train. The vibration velocity and vibration
From Figure 5.13 it can be seen that vibration velocities at the near bridge location
are within the range of (−0.5 ~ + 0.5) 𝑚/𝑠. The magnitude is almost ten times of that at
the open track location. Larger vibration velocity primarily comes from inadequate support
of the hanging crossties. Rail registers the smallest positive peak vibration velocity of 0.15
𝑚/𝑠. Ballast mass vibrates with a slightly higher velocity of 0.2 𝑚/𝑠. The peak vibration
Figure 5.14 clearly demonstrates that the track system accelerations at the near
bridge location are significantly higher than those obtained at the open track location. The
standard gravity constant 𝑔 = 9.8 𝑚/𝑠 M is used as the measurement unit. Recall that the
accelerations at the open track location were all below 1𝑔. While the accelerations at the
near bridge location are all above 10𝑔. In fact , the peak vibration acceleration of rail is
about 15 𝑔 under ACELA Express passenger train. Ballast masses experience larger
accelerations under loading with a peak vibration acceleration of 20𝑔. And the crosstie
exhibits the highest vibration acceleration of up to 40𝑔. These high acceleration values
were also reported in a recent related study by Wilk (2017) on track responses with hanging
ties. According to Wilk (2017), tie-ballast impact forces are observed when an unsupported
tie impacts the ballast during train loading within a short duration and the accelerations range
from about 20 to 40g, depending on the magnitude of the tie-ballast gap. Figure 4.10 shows an
125
example of a tie-ballast gap caused behavior, in which the impact produces a 25 to 30g load
for each pass of a double-train truck (back truck and front truck of next car). Higher
accelerations in the track system can lead to component degradation including ballast
breakdown more easily. In conclusion, the track model prediction results imply a greater
Figure 5.13 Structural Vibration Velocities with Train Passage – Near Bridge.
126
Figure 5.14 Structural Vibration Accelerations with Train Passage – Near Bridge.
track responses. ACELA Express passenger trains moving at 50 mph and 90 mph were
studied using the model to predict and compare responses to those obtained with an
operating speed of 110 mph at the NEC bridge approaches. Figures 5.15 and 5.16 present
the predicted total vertical displacement and vibration responses at the open track and near
bridge locations, respectively. For both locations, the vertical displacements in general do
not increase with the increasing train speed. At the open track location, the vibration
velocities and accelerations increase significantly as the train speed increases; however, at
locations with tie gap, the vibration responses are dominated by the tie gap and are not
127
1.8 1.8 1.8
1 1 1
0 0 0
Ballast
Ballast
Ballast
0 0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Crosstie
Crosstie
Crosstie
0 0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Rail
Rail
Rail
0 0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Time (sec) Time (sec) Time (sec)
Ballast
Ballast
0 0 0
-5 -5 -5
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
5 5 5
Crosstie
Crosstie
Crosstie
0 0 0
-5 -5 -5
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
5 5 5
Rail
Rail
Rail
0 0 0
-5 -5 -5
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Time (sec) Time (sec) Time (sec)
Figure 5.15 Track Vertical Displacements Predicted at Train Speeds of (a) 50 mph, (b) 90
mph and (c) 110 mph at the Open Track Location; Track Vertical Vibration Velocities
Predicted at Train Speeds of (d) 50 mph, (e) 90 mph and (f) 110 mph at the Open Track
Location; Track Vertical Vibration Accelerations Predicted at Train Speeds of (g) 50
mph, (h) 90 mph and (i) 110 mph at the Open Track Location.
128
3 3 3
1 1 1
0.5
0.5 0.5
0
0 0
-0.5
0 1 2 3 4 5 6 7 8 9 10 -0.5 -0.5
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Time (sec)
Time (sec) Time (sec)
Ballast
Ballast
0 0 0
Crosstie
Crosstie
0 0 0
Rail
0 0 0
Ballast
0 0 0
Crosstie
Crosstie
0 0 0
Rail
0
Rail
0 0
-500 -500
0 1 2 3 4 5 6 7 8 9 10 -500
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Time (sec) Time (sec) Time (sec)
Figure 5.16 Track Vertical Displacements Predicted at Train Speeds of (a) 50 mph, (b) 90
mph and (c) 110 mph at the Near Bridge Location; Track Vertical Vibration Velocities
Predicted at Train Speeds of (d) 50 mph, (e) 90 mph and (f) 110 mph at the Near Bridge
Location; Track Vertical Vibration Accelerations Predicted at Train Speeds of (g) 50
mph, (h) 90 mph and (i) 110 mph at the Near Bridge Location.
129
5.6.2 Weight effect
ACELA Express passenger train used previously for model validation with the
locomotive car body weighing 31,550 kg and vehicle car body weighing 45,000 kg is used
again in this section. To quantify the train weight effect on vertical displacement, vibration
velocity, and vibration acceleration responses, the locomotive weight is increased to 52,800
kg and the vehicle car body weight is increased to 57,664 kg, thus simulating a freight train
weight configuration. Note that the purpose of this analysis is to evaluate the train weight
assumptions are made: (i) The train geometry remains the same at both weights for the
passenger and freight conditions; (ii) the bogie is assumed to be 2,400 kg and the wheel is
assumed to be 1,200 kg at both freight train weight and passenger train weight levels; and
Figure 5.17 presents the track responses predicted at the open track location with
different train weights. The vertical responses generated by heavier train weight (freight)
is higher than those generated by lighter train weight (passenger). A locomotive of 31,550
peak displacement of 1.8 mm. Similar trends are found for vibration responses at the open
track location. At the near bridge location (see Figure 5.18), the predicted vertical
displacements due to the heavier train are much higher than those due to the typical
ACELA Express passenger train, however, the vertical vibration velocities and vibration
accelerations only show slight increases for the heavier freight train.
130
1.8 1.8
1.6 1.6
1.4 1.4
Vertical Displacement (mm)
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(a) (b)
0.05 0.05
Ballast
Ballast
0 0
-0.05 -0.05
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0.05 0.05
Crosstie
Crosstie
0 0
-0.05 -0.05
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0.05 0.05
Rail
Rail
0 0
-0.05 -0.05
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(c) (d)
10 10
Ballast
Ballast
0 0
-10 -10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
10 10
Crosstie
Crosstie
0 0
-10 -10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
10 10
Rail
Rail
0 0
-10 -10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(e) (f)
Figure 5.17 Track Vertical Responses for the Heavier Freight Train (a), (c), (e) Compared
to those for ACELA Express Train (b), (d), (f) at the Open Track Location.
131
3 3
2.5 2.5
1 1
0.5 0.5
0 0
-0.5 -0.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(a) (b)
Ballast
0 0
-0.5 -0.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0.5 0.5
Crosstie
Crosstie
0 0
-0.5 -0.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0.5 0.5
Rail
Rail
0 0
-0.5 -0.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(c) (d)
Ballast
0 0
-500 -500
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
500 500
Crosstie
Crosstie
0 0
-500 -500
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
500 500
Rail
Rail
0 0
-500 -500
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(e) (f)
Figure 5.18 Track Vertical Responses for the Heavier Freight Train (a), (c), (e) Compared
to those for ACELA Express Train (b), (d), (f) at the Near Bridge Location.
132
5.7 Summary
with field measured data was presented. The near bridge location at Upland Street was 15
ft. away from the abutment. The open track location was 60 ft. away from the bridge
abutment. Input parameters of the model were carefully chosen so that the model
predictions could match with the field data. Calculated 95% confidence intervals were also
utilized to evaluate the effectiveness of peak responses obtained from the model. For both
locations, the model predictions successfully matched with the field data collected.
This chapter also investigated the model simulation results including reaction
forces, vibration velocities, and vibration accelerations. Analyses show that near bridge
location experiences: (i) larger vertical deformations, especially for crosstie; (ii) larger
impact forces on rail, crosstie, and ballast layer; (iii) larger vibration velocities; and finally,
train speed and train weight on track dynamic track responses. Results indicated that train
speed does not have a significant influence on the vertical displacements predicted at both
the open track and the near bridge locations. The predicted vibration velocities and
accelerations, on the other hand, increased with the train speed increasing at the open track
location, but this pheonomenon was not obvious at the near bridge location. Vertical
displacements predicted under a heavier freight train were higher at both the open track
It is important to note that rail irregularity was not considered in the developed
133
responses. Future studies are recommended to include rail irregularity and its effects on the
134
CHAPTER 6: EVALUATION OF MITIGATION METHODS
effectiveness of several mitigation methods, which are often adopted to improve the
Suggestions on ballasted track designs for newly constructed track transitions as well as
mitigation methods for existing track bridge approaches are provided based on the
evaluation results. To achieve this goal, various combinations of mitigation methods will
be carried out using the validated model. Possible methods include changing of crosstie
6.1 Introduction
There have been several studies proposing in-field mitigation methods for existing
tracks aimed at eliminating the differential movement at railroad track bridge approaches.
Most of the evaluation criteria of proposed mitigation methods highly depend on in-field
performance trends. Field applications of such mitigation methods are usually expensive
and time consuming. Moreover, if the solution conducted is not effective as assumed, it
can result in a waste of resources such as labor and time. Therefore, it is of vital importance
to evaluate these possible mitigation methods for effectiveness potentially with a realistic
track model before any field implementation. This chapter intends to utilize the vehicle-
track-bridge model to conduct parametric studies and evaluate the effectiveness of possible
mitigation methods.
From the structural perspective, there are four major components in the track
system, namely rail, crosstie, ballast, and subgrade. Usually, the rail beam properties are
135
consistent across the track transition zone as the rail beam spreads over a long distance in
the field. Crossties, ballast, subgrade (open track side) and pads between the major
components, on the contrary, are adjustable along the train moving direction. Therefore,
the mitigation plans are focused on modifying the properties of aforementioned track
components for providing a smoother transition between the track bed and bridge deck.
To achieve this ultimate goal, individual effects of rail pad properties, crosstie
properties, and ballast and subgrade properties on the bridge approach dynamic responses
need to be assessed. Based on such a sensitivity study, two mitigation plans are proposed
to be studied in this chapter. The first one is installation of under tie pads at near bridge
locations, and it can be achieved by modifying the damping ratio and stiffness inputs in the
layer on an open bridge deck. This can be simulated by adding a layer of ballast mass and
The effects of changing selected track properties on the dynamic responses have
been studied with the Train-Track-Bridge model. The selected track parameters include:
(1) rail pad stiffness, (2) crosstie spacing, (3) ballast stiffness, and (4) subgrade stiffness.
Table 6.1 lists the default values for these track parameters in the track model. Sections
136
Table 6.1 Default Parameters in the Track Model
Within a typical rail pad stiffness range, six values (between 10 𝑡𝑜 1000 𝑀𝑁 ⁄𝑚 /
𝑚) were selected to run the Train-Track-Bridge model at both near bridge and open track
locations. Figures 6.1 and 6.2 show the simulated effects of the rail pad stiffness on the
magnitude of peak transient deformation and peak reaction force between rail and crosstie
at both locations. The results show a clear negative correlation between rail pad stiffness
and vertical deformation, and a clear positive correlation between rail pad stiffness and
rail-tie reaction force. From the perspective of minimizing vertical displacement and
lowering the rail-tie reaction force, the current train-track-bridge model appears to yield an
optimum rail pad stiffness value. It can be seen that when the rail pad stiffness is lower
than 100 𝑀𝑁⁄𝑚 /𝑚, the transient vertical deformation increases rapidly with decreasing
rail pad stiffness, at both near bridge and open track locations. The simulation results also
indicate that at open track location, rail pad stiffness of approximately 100 𝑀𝑁⁄𝑚 /𝑚
could be a turning point for the magnitude of rail-tie reaction force, after which, the
increase of rail-tie reaction force gradually slows down with increase of rail pad stiffness.
While at near bridge location, the rail-tie reaction force continues to increase at a high rate
with increasing rail pad stiffness until stiffness reaches 500 𝑀𝑁⁄𝑚 /𝑚.
137
Figure 6.1 Effect of Rail-Pad Stiffness on Displacement
138
6.2.2 Effect of crosstie spacing
The sensitivity study of crosstie spacing is carried out via the numerical Train-
Track-Bridge model. Crosstie spacing refers to the distance between two adjacent crossties.
In the field testing and original model setup, the distance between instrumented concrete
crossties has been fixed at 0.6096 m (2 ft.). In practice, it is anticipated that larger spacing
leads to higher deformation as the track stiffness could be lowered. From the simulation
results, it is confirmed that the vertical deformation increases with increasing crosstie
spacing (Figure 6.3). The near bridge location results indicate that by decreasing the
crosstie spacing from 0.6 m to 0.4 m, the vertical displacement decreases by around 10%.
The open track location results indicate a more significant effect from the same amount of
139
The magnitude of crosstie spacing also affects greatly the rail-tie reaction force.
The relationship of crosstie spacing and rail-tie reaction force at both open track and near
bridge locations is presented in Figure 6.4. The rail-tie reaction force increases when the
spacing is increased from 0.4 m to 0.6096 m. Specifically, the maximum value of rail-tie
whereas the maximum value is only 28 kN with a crosstie spacing of 0.4 m. Similarly, at
near bridge location, the value of rail-tie reaction force can be as high as 92 kN with the
current crosstie spacing value of 0.6096 m, while it decreases to 79 kN when the crosstie
crosstie spacing in a track system design to mitigate the high deformation and reaction
140
6.2.3 Effect of ballast stiffness
Five ballast stiffness values were selected to conduct sensitivity study through the
and the highest value is 1000 𝑀𝑁⁄𝑚 /𝑚. Figures 6.5 and 6.6 illustrate the effect of ballast
stiffness on vertical deformation and rail-tie reaction force, respectively. It is observed that
ballast stiffness has a major impact on the total track system displacement. By increasing
ballast stiffness, the maximum vertical deformation in the track system decreases from 4.3
mm to 1.3 mm at open track location, and from 4.3 mm to 2.3 mm at near bridge location.
Note that the ballast stiffness has a more significant effect on open track location, it is
because the near bridge location encounters a gap between crosstie and ballast layer, thus
the decrease of ballast stiffness will not have a huge impact on the original “larger”
deformation.
Similarly, in Figure 6.6, it can be seen that the rail-tie reaction force increases when
the ballast stiffness increases. Maximum incurred reaction force at open track location is
increases to 1000 𝑀𝑁⁄𝑚 /𝑚 , the maximum calculated reaction force at open track
location is 42 kN. At near bridge location, the rail-tie reaction force increases from 34 kN
to 280 kN as the ballast stiffness increases. Interestingly, even though the ballast stiffness
has a smaller impact on deformation at near bridge location, it has a much larger influence
transition design process to generate an optimal track system. If the ballast stiffness is too
small, the track system yields a high deformation, which is detrimental to the long-term
141
track geometry; on the other hand, if the ballast stiffness is too large, the track system
withstands high forces on track components, which is harmful to the longevity of structure.
142
6.2.4 Effect of subgrade stiffness
The effect of subgrade stiffness was studied through the Train-Track-Bridge model
by altering the subgrade stiffness parameter in the simulations. The simulations were
conducted for four subgrade stiffness input values, including 10, 50, 100, and 1000
𝑀𝑁⁄𝑚 /𝑚.
The calculated track structural responses, i.e., maximum vertical deformation and
maximum rail-tie reaction force, are graphed in Figures 6.7 and 6.8, respectively. The
vertical deformations are reduced rapidly with increasing subgrade stiffness until
100 𝑀𝑁⁄𝑚 /𝑚, followed by a slower rate of decrease. In a similar manner, the reaction
force increases rapidly with increasing subgrade stiffness until 100 𝑀𝑁⁄𝑚 /𝑚, and slows
at the two examined locations. At open track location, the maximum calculated vertical
𝑚, which is about 3 times the original deformation (0.9 mm). At near bridge location, the
increase in deformation is approximately 2.5 mm, which is slightly higher than the original
deformation (2.2 mm). The effects of subgrade stiffness on rail-tie reaction force are quite
In addition to the previous findings, it is important to note that the effect of subgrade
stiffness on vertical deformation is noticeably larger than that on reaction force. This
indicates that track system deformation is more sensitive to subgrade stiffness. And the
rail-tie reaction force in the system is not affected significantly by the subgrade stiffness.
143
Figure 6.7 Effect of Subgrade Stiffness on Vertical Displacement
144
6.3 Proposed Mitigation Plan I – Installation of Ballast on Bridge Deck
Traditionally railroad bridges are open deck, which means there are no ballast
materials placed on the bridge deck to support the track system. Researchers have
suggested converting open-deck bridges to ballasted deck ones to mitigate track transition
problems (Hyslip et al. 2009). According to the previous studies, the alternatives of
effective mitigation plans should ease the transition from approach embankments to the
fixed bridge structure. It is important to test whether installation of ballast layer on bridge
deck could provide such an easier transition. Hence, converting an open deck bridge to a
ballasted deck bridge was selected as one of the mitigation plans for reducing differential
movement at bridge approaches. Figure 6.9 illustrates the original track structure on bridge
The proposed plan is to install the same quality of ballast materials on the bridge.
Therefore, all the parameters related to ballast (ballast density, ballast stiffness, and ballast
damping coefficient) are chosen the same as those of the ballast on the embankment side.
145
(a) Current Open Deck Bridge
Figure 6.9 Illustration of Converting Open Deck Bridge to Ballasted Deck Bridge.
conducting a parametric study with the Train-Track-Bridge model. As shown in Figure 6.9,
ballast masses are added at the bridge side. The gap between tie and ballast at bridge
approach is approximately 1.5 mm. The spring stiffness underneath the ballast mass is still
the same as the previous stiffness on top of bridge deck, which is 100 times higher than the
subgrade stiffness on the embankment side to accommodate the effect of stiff bridge. The
observation point is at the 10th crosstie away from the bridge abutment. Track responses at
146
this observation point were calculated by the model and analyzed for before and after
converting an open deck bridge to ballast deck bridge. The reaction forces predicted in the
(a) (b)
(c) (d)
Figure 6.10 Calculated Reaction Forces (a) Rail-Tie before (b) Rail-Tie after (c) Tie-
Ballast before (d) Tie-Ballast after Converting to Ballasted Deck.
From Figure 6.10, it can be seen that installation of ballast layer on bridge deck can
slightly reduce the reaction forces in the track structure. The peak force between rail and
147
crosstie decreases from over 90 kN to 80 kN; and the peak extensive force between crosstie
and ballast mass decreases from 50 kN to 30 kN. Therefore, even though the gap still exists
at near bridge locations, the resultant impact loads between track components can be
Figure 6.11 demonstrates the calculated transient displacement time history before
and after converting an open deck to a ballasted deck bridge. Interestingly, the installation
of ballast materials on bridge deck does not help to mitigate the high vertical deformations
at near bridge location. On the contrary, the displacements are slightly increased as a result
of the proposed mitigation method. Therefore, ballast on bridge providing a slightly softer
track structure at bridge side will not benefit reducing deformations at near bridge location
where tie gaps may exist. An additional effort may be needed to handle such a problem.
(a) (b)
Figure 6.11 Calculated Vertical Deformation (a) before and (b) after Installation of
Ballast on Bridge Deck.
Besides the transient deformation and force time histories, ballast and crosstie
vibration characteristics are also analyzed using the track model. Figure 6.12 (a) and (b)
148
show the ballast vibration velocities before and after converting to a ballasted deck bridge.
It is noted that peak ballast vibration velocity is greatly reduced from 0.22 m/s to 0.15 m/s.
Moreover, crosstie vibration velocity after the proposed mitigation method has decreased
from 0.25 m/s to 0.2 m/s as shown in Figure 6.12 (c) and (d). Similarly, the ballast and
crosstie vibration accelerations are also calculated and graphed in Figure 6.12 (e) to (h).
Before any mitigation methods, the peak ballast vibration acceleration can be as high as 20
g (gravity coefficient), while the peak crosstie vibration acceleration can reach 30 g, as a
result of hanging tie. Installation of ballast on a bridge deck can help reduce the ballast
impact loads indicate that ballasted deck bridges can better maintain the track geometry
performance indexes of the track are improved by the proposed mitigation method –
converting an open deck bridge to ballasted deck. However, installation of ballast on bridge
deck will not help to mitigate the existing gap and the gap related large deformation and
differential movement issue at near bridge location. Hence, additional methods, for
example, stone blowing at hanging tie locations are needed to treat the problem. Such a
stone blowing application was carried out successfully at the Madison Street bridge
149
(a) (b)
(c) (d)
Figure 6.12 Calculated Dynamic Responses (a), (c), (e), (g) before and (b), (d), (f), (h)
after Converting Open Deck Bridge to Ballasted Deck Bridge.
150
Figure 6.12 (Continued).
(e) (f)
(g) (h)
Under Tie Pads (UTPs) are essentially elastic pads placed underneath the crossties
on top of the ballast. The UTPs have been found to reduce the level of vibration caused in
the ballast layers under train loading, and also increase the damping effect of the track
substructure. Moreover, UTPs can reduce the influence of varying track stiffness on the
151
wheel/rail contact force, and as a result distribute the load of the train over a wider area
within the ballast layer. Several successful research and implementation efforts involving
the use of UTPs were discussed in Chapter 2 of this dissertation. Owing to their ability to
facilitate better load transfer onto the ballast layer, UTPs were selected as one of the
remedial measures in the Northeast Corridor field study (Tutumluer et al. 2016). Figure
6.13 shows a schematic layout of the bridge approaches and the locations where individual
Figure 6.13 Schematic Layout of Selected Bridge Approaches and Relative Locations of
Remedial Measures (Tutumluer et al. 2016).
A block of 30 crossties with UTPs attached to bottom of the ties were installed on
the Upland Street Bridge south approach. Tracks No. 1 and 4 were kept active during the
installation but No. 2 and No. 3 were out of service to complete the installation. A pre-
constructed concrete tie track panel was brought on site for installation. The installation
was performed by the Amtrak work crews and involved removing the old track, installing
the new track panel, unloading ballast and compacting the track substructure, surfacing and
aligning the track, and finally thermite welding the rail joints. Work was completed in a
weekend by midnight on Saturday 30 August 2014, well within the track outage limits.
152
According to the manufacturer Pandrol-CDM Track / Novitec, the installed UTPs
were made from polyurethane elastomers manufactured in France from recycled materials
(ground tires), with a shore hardness of 65 to 75. These pads had on average 10% void
content to allow better ballast fit. Based on track geometry records at the Upland Street
south approach (Track No. 2), it was decided that installing 30 new ties with UTPs would
UTPs with desired engineering properties were shipped to the Amtrak Wilmington Yard
facility. To ensure full adherence of the pads to the bottom of the newly fabricated concrete
ties, the surface was prepared to eliminate all fines or dust. A special adhesive was placed
on the bottom of the crosstie, then the pads were glued and set to dry for 8 to 12 hours. The
tie bottoms were placed mirroring each other with an insulation material (thin carpet) in
order to use the weight of the tie to press down pad and concrete binding. The elastomers
were cut and installed with half an inch offset from the edge of the tie.
The first step in the installation process was cutting the rail at four locations to
remove the 30-crosstie panel from Track No. 2. The new track panel was approximately
60 ft. in length and constructed with 30 concrete ties and 136# RE rail. It was placed on
the east side of the bridge ensuring that 24-in. spacing was standardized between ties. A
120-ton capacity Kirow crane was used to remove the old track panel. Once the space was
cleared, a front-end loader excavated the old ballast from Track No. 2. Ballast was removed
to a depth of 27 inches from the top of rail (TOR), coinciding with the layer of asphalt that
had previously been placed under the tracks. A small amount of ballast was left as a layer
to accommodate the new panel. Next, the new track panel was installed, the rail was cut in
153
order to have staggered welds and finally, the joint bars were fastened. A photograph of
Figure 6.14 Installation of a New Track Panel Having 30 New Concrete Crossties with
Under Tie Pads at the Upland Street Bridge Approach Site near Chester PA.
To test the effectiveness of installing under tie pads, track geometry data and
transient track response data were analyzed. Figure 6.15 shows the space curve and running
roughness data for the Upland Street south approach (Track 2) immediately before and
after installation of the track panel with the UTP mounted crossties. As shown in the figure,
installation of the UTPs in late August 2014 resulted in an immediate increase in the
running roughness and is represented by an upward shift in the space curve. This can be
attributed to uneven resurfacing of the track immediately after installation of the track
panels. However, note that the space curve attains a stable configuration within two weeks
after installation of the track panel with UTP mounted ties. A reduction in the running
154
roughness was observed in December 2014. No significant shift in the space curve, or
increase in the roughness is observed till November 2015. This indicates that the track
attained a “stable” configuration after an initial “shakedown,” and installation of the UTPs
80
0.5
Space Curve
60
0.0
Roughness
40
-0.5
20
-1.0 0
6/1/14 8/1/14 10/1/14 12/1/14 2/1/15 4/1/15 6/1/15 8/1/15 10/1/15 12/1/15
Figure 6.15 Space Curve and Running Roughness Data for the Upland Street Bridge
Approach Showing the Effect of UTP Installation.
A separate instrumentation effort was carried out in August 2015 to measure the
transient response of the track panel under train loading. Note that this instrumentation
effort was carried out eleven (11) months after installation of the track panel with UTP
mounted ties, and hence can be used as an indicator of long-term performance of the UTPs.
Figure 6.16 shows a schematic of the instrumentation layer to measure the transient
response under train loading for this particular bridge approach. As shown in the figure,
the instrumentation was mounted on the 10th and 11th ties from the bridge abutment.
155
Considering a 24-inch tie spacing, these ties are located at distances of approximately 20
Note that the instrumentation effort for UTPs are different from the other locations
mentioned in previous chapters as the transient deformation herein only measures the
ballast layer deformation. Figure 6.17 presents load and displacement time histories
measured at this bridge approach during the passage of an Acela Express train. As shown
in the figure, peak transient displacements of approximately 1.5 mm were measured for
both Tie No. 10 and Tie No. 11. Considering that this measurement was done almost a year
after installation of the track panel with UTP mounted ties, the low peak transient
adequate support conditions underneath both ties existed even one year after the track panel
installation.
Figure 6.17 Force and Displacement Time Histories Measured at the Upland Street
Bridge Approach.
To evaluate the effectiveness of installing Under Tie Pads (UTPs) on the dynamic
behavior from a mathematical perspective, a parametric study was also conducted using
the Train-Track-Bridge model. At first, the original structural responses at Tie #10 (away
from abutment) were calculated as a baseline. The tie gap value is the same as used in the
previous analysis, which is approximately 1.5 mm. Next, some track parameters were
modified to simulate adding 30 UTPs at near bridge location. Newly calculated track
structural responses were compared with the original ones before installation of UTPs.
157
Several assumptions were made for estimating the condition of track system with
UTPs. Under Tie Pads have been previously studied to show their effects on the track
system properties by decreasing the stiffness and increasing the damping ratio. Therefore,
it is assumed that at the locations with UTP installation, the ballast stiffness will decrease
from 120 𝑀𝑁⁄𝑚 /𝑚 to 50 𝑀𝑁⁄𝑚 /𝑚 and ballast damping ratio will increase from 0.082
𝑀𝑁. 𝑠⁄𝑚 /𝑚 to 0.15 𝑀𝑁. 𝑠⁄𝑚 /𝑚. Moreover, since the thickness of UTPs are usually
over 1 mm, it is assumed that the gap between crosstie and ballast layer will be closed.
simulations were carried out. The observation point was set at 10 crossties away from the
bridge abutment, the same as in the field condition. Figure 6.18 shows the reaction forces
that occurred in the track system before and after UTP installation. Note that since the
thickness of under tie pads would decrease the amount of gap between tie and ballast, the
158
(a) (b)
(c) (d)
Figure 6.18 Calculated Reaction Forces (a) Rail-Tie before (b) Rail-Tie after (c) Tie-
Ballast before (d) Tie-Ballast after Under Tie Pads (UTPs) Installation.
Figure 6.19 shows the vertical deformation at top of rail, crosstie, and ballast layer
before and after UTP installation at the same location. Before conducting any mitigation
methods, the top of rail vertical deformation upon train passage can be as high as 2.5 mm,
majorly due to the hanging crosstie at near bridge location. With installation of the 30 UTPs
at these problematic locations, the total vertical deformation is now lowered to 2 mm, even
159
(a) (b)
Figure 6.19 Calculated Vertical Deformation (a) before and (b) after Under Tie Pad
(UTP) Installation.
Taking a closer look into the dynamic behavior of the track system, velocities and
accelerations of the track components are calculated in the model. Figure 6.20 (a) and (b)
illustrates the ballast velocities underneath the 10th crosstie from bridge abutment before
and after UTP installation. Note that the axis limit differs in the two subplots in order to
show the velocity after UTP clearly. The maximum ballast velocity under train passage
decreases from 0.2 m/s to 0.02 m/s. This is mainly attributed to the (i) mitigation of hanging
crosstie condition and (ii) increased damping coefficient due to UTP installation. Figure
6.20 (c) and (d) demonstrates the velocity time histories of the observed crosstie. The
vibration velocity of crosstie can also be controlled via increasing damping coefficient and
mitigating gap between crosstie and ballast. Figure 6.20 (e) to (h) show the ballast and
derivative of velocity, it follows the similar trend of the vibration velocity. Simulation
results indicate that installation of UTPs can potentially help reduce the ballast and crosstie
160
accelerations tremendously from 20 g to 0.2 g (g is gravity constant). Track vibration
(a) (b)
(c) (d)
Figure 6.20 Calculated Dynamic Responses (a), (c), (e), (g) before and (b), (d), (f), (h)
after Under Tie Pad (UTP) Installation.
161
Figure 6.20 (Continued).
(e) (f)
(g) (h)
In this section, the validated train-track-bridge model is adopted to study the train
ride quality. Ride quality, also refers to as riding comfort, is a key factor in the designs of
high-speed passenger trains. According to research studies, the riding comfort can usually
be evaluated by the vertical acceleration of the car body and this value should be limited
to a certain maximum to ensure adequate riding comfort (Youcef et al., 2013; Ziyaeifar,
2008; Kargarnovin et al., 2005; Yean and Yeong, 2003; Jong et al., 1999). To study the
162
train ride quality at bridge approaches, ACELA Express passenger train is simulated to
move from the open track onto the bridge deck. The time history of car body acceleration
at the regular operating speed of 110 mph is graphed. In addition, four different speeds
from 22 mph to 200 mph have also been studied to investigate the speed effect. The number
of ties on track and bridge was limited 500 as in the previous modeling cases. To eliminate
any influence of boundary conditions on car body acceleration, the number of ties on the
open track side was increased to 600, 800, 1000, 1200, and 1500 for the 22, 67, 110, 156,
and 200 mph cases, respectively. In accordance, the simulation results obtained are graphed
in Figure 6.21. Because track irregularity is not considered in the model, car body
acceleration at open track side is almost negligible. When the train arrives at a bridge
approach where there are gaps underneath the crossties, the car body acceleration
immediately increases. Similarly, when the train arrives at a transition point between the
bridge deck and open track, the car body acceleration has another significant jump and then
gradually decreases. Figure 6.22 shows the peak car body acceleration varying with the
train speed. Results show that with the increase of train speed, the peak car body
acceleration also increases. When the train operates at 110 mph, approximate peak car body
acceleration at the near bridge location with hanging ties is around 0.11 m/s/s. When the
operating speed increases to 200 mph (high-speed train), the car body acceleration becomes
0.15 m/s/s. Note that in the field the anticipated car body acceleration will be higher due to
track irregularity. At this stage, the train-track-bridge model has been shown to indicate
riding comfort with idealized examples. Once the track irregularity is introduced, the
vertical car body acceleration can be used to measure the train ride quality quantitatively.
163
0.15 0.15
0 0
-0.05 -0.05
-0.1 -0.1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(a) (b)
0.15 0.15
0 0
-0.05 -0.05
-0.1 -0.1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec) Time (sec)
(c) (d)
0.15
Embankment Bridge
0.1
Car Body Acceleration (m/s/s)
0.05
-0.05
-0.1
(e)
Figure 6.21 Car Body Acceleration at Train Speeds of (a) 10 m/s (~22 mph), (b) 30 m/s
(~67 mph), (c) 50 m/s (~110 mph), (d) 70 m/s (~156 mph), and (e) 90 m/s (~200 mph).
164
0.16
Acceleratopm (m/s/s)
0.12
0.08
0.04
0
0 50 100 150 200 250
Train speed (mph)
6.6 Summary
This chapter was intended to provide some insights for future design and
maintenance activities. First, parametric studies were conducted using the Train-Track-
Bridge model by changin rail pad stiffness, ballast stiffness, subgrade stiffness, and crosstie
spacing. Simulation results show that increasing rail pad stiffness, ballast stiffness, and
subgrade stiffness at near bridge approach is effective to decrease the level of settlement at
track bridge approaches. However, increasing the system stiffness could imply higher
should be selected for ballasted track design to gain optimal system responses. In addition,
results from the model indicate that crosstie spacing is crucial to both vertical deflection
and reaction force. Smaller spacing than typical 2-ft. spacing can reduce the vertical
deformation as well as reaction force between track components. Hence, constructing the
165
bridge approach with gradually decreased crosstie spacing could help to prevent high
Next, this chapter presented details on the selection of two mitigation methods for
the problem of differential movement at the instrumented railroad track transition sites at
the Amtrak Northeast Corridor lines. The studied remedial measures were (1) installation
of ballast layer on open deck bridge and (2) installation of a track panel having new
crossties with Under Tie Pads (UTPs). They were selected based on the review of published
literature and site inspection for effectiveness at the Amtrak Northeast Corridor. The model
simulation results indicated that installation of ballast layer on bridge deck could help
mitigate differential movement problems and reduce impact forces between track
components, and vibration velocities and accelerations of crosstie and ballast. Therefore,
it is a good remedial application to maintain the current good track performance and slow
down track geometry deterioration in the future. However, it will not help to mitigate the
high vertical deformation that may occur at near bridge locations. If the track in the field
is already experiencing hanging tie issues, more effective remedial measures, such as stone
Field instrumentation data were collected after installation of a new track panel
having 30 concrete ties with UTPs at the Upland Street bridge approach on Amtrak’s
Northeast Corridor line. The field data showed that after one year of UTP installation, the
original shakedown process of newly added ballast was fully completed. Transient
displacement data indicated good crosstie and ballast contact at the near bridge location.
To further examine the dynamic behavior such as velocities and accelerations of track
166
proved the installation of UTPs to be an effective remedial measure because UTP installed
ties exhibited reduced (1) impact loads and reaction forces, (2) vertical deformations, and
Lastly, the proposed train-track-bridge model was adopted to study ride quality.
Vertical car body acceleration was chosen as an indicator for high-speed train riding
comfort. Results show that at open track location, the car body acceleration was negligible;
while at near bridge location, car body acceleration increased at two locations, the starting
In this chapter, the model simulation results provided some insights into possible
solutions to the differential movement problem at track transition zones, i.e. bridge
approaches specifically. The developed vehicle-track-bridge model has been proven for its
scenarios to help gain deeper understanding of the track system dynamic responses at
bridge approaches. Further, the studied mitigation methods through the model simulaions
167
CHAPTER 7: CONCLUSIONS AND RECOMMENDATIONS
effectively simulate and quantify the dynamic response and performance behavior of track
structures at the transition zone of a bridge approach under moving train loads. In
accordance, the research work undertaken focused on the statistical field data analysis,
of the developed track model in order to provide improved bridge approach designs and
track maintenance and rehabilitation practices. The objective was to better understand
governing mechanisms of bridge approach problems that occur near bridge abutments and
utilize the train-track-bridge model to help with designing new bridge approaches and
selecting mitigation methods. This chapter summarizes the main research findings and
contributions of this thesis study and provides suggestions for future work.
speed passenger lines near Chester, Pennsylvania, including calculated means, standard
deviations, and 95% confidence intervals. Chapter 4 presented the developed ballasted
train-track-bridge model and solving scheme for both open track (linear) and near bridge
(nonlinear) locations. Chapter 5 presented the validation of the simulation results from the
developed ballasted train-track-bridge with the field data. The validation results showed
that the deformations obtained from the track model successfully matched with the
deformation data collected from the field instrumentation. Chapter 6 discussed various
168
mitigation methods applicable to track bridge approaches including changing rail pad
stiffness, ballast stiffness, subgrade stiffness, and crosstie spacing. Furthermore, Chapter 6
presented details on the selection of two mitigation methods for remedying the problem of
differential movement at the instrumented railroad track transition sites at the Amtrak’s
Northeast Corridor, namely, (1) converting an open deck bridge to ballasted deck bridge
and (2) installation of a new track panel having concrete ties with Under Tie Pads (UTPs).
Based on the study, the following findings are drawn concerning the field
increasing trend with time, while the vertical wheel loads were relatively stable,
dynamic responses measured at open track and near bridge locations; much higher
vertical deformations and vibrations velocities and accelrations were taking place
- Tie gaps existed at both open track and near bridge locations. The amount of tie
gap at near bridge location could be as high as 3.9 mm at the investigated Upland
Street bridge approach site. Whereas, the amount of tie gap at open track location,
60 ft away from the Upland Street bridge abutment, was approximately 0.2-0.3 mm.
nonlinear spring between crosstie and ballast masses in the novel train-track-bridge
model.
169
- Nonlinear properties could be properly accounted for in the developed track model
scheme.
- Track model predictions indicated that at open track locations, each crosstie took
approximately 30%-32% of the train wheel load. On the contrast, crossties might
properly increasing rail pad stiffness, ballast stiffness, and subgrade stiffness.
- Decreasing crosstie spacing was shown to reduce the vertical deformation as well
as reaction forces between track components, which could prevent high dynamic
- Installation of a ballast layer on an open deck bridge could help to reduce impact
would not help to mitigate the high vertical deformations often occurred at near
bridge locations. If the track in field already suffered from major hanging tie issues,
such as existing impact forces and ballast-tie gaps under train loading, additional
problem.
- Field instrumentation data indicated good crosstie and ballast contact at near bridge
locations almost one year after the installation of a new track panel having 30
concrete crossties with UTPs. Numerical simulation of UTP installation also proves
170
installation of UTPs to be effective remedial measures because UTP installed ties
exhibited reduced (1) impact loads and reaction forces, (2) vertical deformations,
and (3) vibration velocity and acceleration responses due to train loading.
that could be used to test the effectiveness of proposed mitigation methods for solving
transition zone problems as well as improve future track designs at bridge approaches.
Many questions remain open regarding more detailed structural analysis and field
performance prediction. There are several important directions for future research.
bridge model to include lateral forces and derailment potential analyses. Upon
improvement of the existing model, one can also take track irregularity into consideration
and establish tamping and other track maintenance schedules for maintaining proper track
Another important question concerns optimal choices of track input properties for
overall track design and maintenance guidance. For example, if necessary, the track design
outputs given a certain track condition, such as train speed, subgrade modulus, etc.
Finally, it would be useful to collect field geometry data over a long period of time
and combine it with the numerical model for the transient force and displacement field data
collection. The combined permanent and transient deformation data can be analyzed using
171
machine learning techniques to predict important indexes of track structure, such as
172
REFERENCES
[1] Banimahd, M., Woodward, P. K., Kennedy, J., Medero, G. M. Behavior of train–
[2] Basu, D., Kameswara Rao., N. S. V. Analytical solutions for Euler–Bernoulli beam
[3] Bian, X., Chen, Y., Hu, T. Numerical simulation of high-speed train induced
ground vibrations using 2.5 D finite element approach. Science in China Series G:
[4] Bilow, D., Li, D. Concrete slab track test on the high tonnage loop at the
Conference; 2005.
[5] Biondi, B., Muscolino, G., Sofi, A. A substructure approach for the dynamic
2281.
[6] Boler, H. (2012). On the shear strength of polyurethane coated railroad ballast.
[7] Boler, H., Mishra, D., Hou, W., & Tutumluer, E. (2018). Understanding track
[8] Boler, H., Mishra, D., Tutumluer, E., Chrismer, S., & Hyslip, J. P. (2019). Stone
173
railroad bridge approaches. Proceedings of the Institution of Mechanical Engineers,
[9] Briaud, J. L., James, R. W., & Hoffman, S. B. (1997). Settlement of bridge
approaches: (the bump at the end of the bridge) (Vol. 234). Transportation Research
Board.
[10] Briaud, J. L., Nicks, J. E., Smith, B. J. The bump at the end of the railway bridge,
[11] Cantero, D., Arvidsson, T., OBrien, E., & Karoumi, R. (2016). Train–track–bridge
12(9), 1051-1064.
[12] Chang, C. S., Selig, E. T., Adegoke, C. W. Geotrack model for railroad truck
1201-1218.
[13] Chawla, S., & Shahu, J. T. (2016). Reinforcement and mud-pumping benefits of
[14] Chrismer, S. Track surfacing with conventional tamping and stone injection.
[15] Coelho, B., Priest, J., Holscher, P., Powrie, W. Monitoring of transition zones in
174
[16] Coelho, B., Hölscher, P., Priest, J., Powrie, W., & Barends, F. (2011). An
Mechanical Engineers, Part F: Journal of Rail and Rapid Transit, 225(2), 129-139.
[17] Czyczula, W., Koziol, P., Blaszkiewicz, D. On the Equivalence between Static and
[18] Davis, S. C., Boundy, R. Transportation Energy Data Book. Oak Ridge National
Laboratory; 2018.
[19] Dersch, M. S., Tutumluer, E., Peeler, C. T., & Bower, D. K. (2010). Polyurethane
coating of railroad ballast aggregate for improved performance. In 2010 Joint Rail
[20] De Beer, M., Horak, E., Visser, A. T. The multidepth deflectometer (MDD) system
[21] Feng, H. (2011). 3D-models of railway track for dynamic analysis. Master Thesis.
[22] FIMOR (2015). Fimor products for track superstructure. In Railway Engineering
[23] Frohling, R. D., Scheffel, H., Ebersöhn, W. The vertical dynamic response of a rail
vehicle caused by track stiffness variations along the track. Vehicle System
[24] Giner, I. G., López Pita, A., Vieira Chaves, E. W., & Rivas Álvarez, A. M. (2012).
175
Proceedings of the Institution of Civil Engineers-Transport (Vol. 165, No. 1, pp.
[25] Gräbe, P. J., Shaw, F. J. Design life prediction of a heavy haul track foundation.
[26] Gräbe, P. J., Mtshotana, B.F, Sebati, M.M., & Thunemann, E.Q. (2015). The effects
[27] Hayano, K., Ishii, K., & Muramoto, K. (2013). Effects of ballast thickness and tie-
report WG4.
[28] Hayano,K., Hoshi, S., & Kumara, J.J. (2013). Soil mass density test on railway
[29] Holm, G., Andréasson, B., Bengtsson, P. E., Bodare, A., Eriksson, H. Mitigation of
track and ground vibrations by high speed trains at Ledsgard, Sweden. Swedish
[30] Huang, H., Shen, S., Tutumluer, E. Sandwich model to evaluate railroad asphalt
176
[31] Huang, H., Gao, Y., Stoffels, S. Fully coupled three-dimensional train-track-soil
[32] Hyslip, J. P., Li, D., McDaniel, C. R. Railway bridge transition case study. In
[33] Indraratna. B., Sajjad, M., Ngo, T., el at. Improved performance of ballasted tracks
[34] Insa, R., Salvador, P., Inarejos, J., & Roda, A. (2012). Analysis of the influence of
[35] Johansson, A., Nielsen, J. C., Bolmsvik, R., Karlström, A., & Lundén, R. (2008).
1479-1487.
[36] Jong, D. Y., Yeong, B. Y. and Shyh, R. K. Impact response of high speed rail
bridges and riding comfort of rail cars. Engineering Structures 1999; 21(9): 836-
844.
177
[38] Judge, A., Ho, C., Huang, H., Hyslip, J., & Gao, Y. (2018). Field Investigation and
[39] Kaewunruen, S., & Tang, T. (2019). Idealisations of Dynamic Modelling for
[41] Kalker, J. J. Discretely supported rails subjected to transient loads. Vehicle System
[42] Kargarnovin, M.H., Younesian D., Thompson, D. and Jones, C. Ride comfort of
high-speed trains travelling over railway bridges. Vehicle System Dynamics 2005;
43(3): 173-197.
[43] Keene, A., & Edil, T. B. (2012). Mitigating Ballast Fouling Impact and Enhancing
Rail Freight Capacity. National Center for Freight & Infrastructure Research &
[44] Kennedy, J., Woodward, P. K., Medero, G., & Banimahd, M. (2013). Reducing
[45] Kerr, Arnold D., and Brian E. Moroney. "Track transition problems and remedies."
[46] Kerr, A., Bathurst, L. (2001). Upgrading track transitions for high-speed service.
178
[47] Koch, E., Hudacsek, P. 3D Dynamic Modeling of Transition Zones. World
prediction method for urban railway vibration. Soil Dynamics and Earthquake
[49] Kouroussis, G., Vogiatzis, K., Connolly, D. Modelling the Environmental Effects
[52] Lei, X., Noda, N. A. Analyses of dynamic response of vehicle and track coupling
system with random irregularity of track vertical profile. Journal of Sound and
[53] Lei, X., & Rose, J. G. (2008). Track vibration analysis for railways with mixed
[54] Le Pen, L., Watson, G., Hudson, A., & Powrie, W. (2018). Behaviour of under
Institution of Mechanical Engineers, Part F: Journal of rail and rapid transit, 232(4),
1049-1063.
179
[55] Li, X., Ekh, M., Nielsen, J. Three-dimensional modelling of differential railway
track settlement using a cycle domain constitutive model. International Journal for
[56] Li, D., & Maal, L. (2015). Heavy axle load revenue service bridge approach
[57] Li, D., Selig, E. T. Method for railroad track foundation design I: Development.
[58] Li, D., Davis, D. Transition of railroad bridge approaches. Journal of Geotechnical
[59] Ling, L., Xiao, X., Xiong, J., el at. A 3D model for coupling dynamics analysis of
15(12): 964-983.
[62] Mise, K., Kunii, S. A theory for the forced vibrations of a railway bridge under the
180
[63] Mishra, D., Tutumluer, E., Stark, T. D., Hyslip, J. P., Chrismer, S. M., Tomas, M.
13(11): 814-824.
[64] Mishra, D., Tutumluer, E., Boler, H., Hyslip, J., Sussmann, T. Railroad track
Research Record: Journal of the Transportation Research Board 2014; (2448): 105-
114.
[65] Mishra, D., Boler, H., Tutumluer, E., Hyslip, J. P. Effectiveness of Chemical
[66] Moale, C., Wilk, S., Smith, D., el at. Design and Performance of Three Remediated
[68] Muramoto, K., Nakamura, T., Sakurai, T. A study of the effect of track irregularity
prevention methods for the transition zone between different track structures.
181
[70] Nicks, J. The bump at the end of the railway bridge. PhD Thesis, Texas A&M
state-dependent track properties. Journal of sound and vibration 2004; 275(3): 515-
532.
[73] Paixão, A., Fortunato, E., Calçada, R. Design and construction of backfills for
Engineers, Part F: Journal of Rail and Rapid Transit 2015; 229(1): 58-70.
[74] Paixão, A., Alves Ribeiro, C., Pinto, N., Fortunato, E., & Calçada, R. (2015). On
[76] Pen, L.L., Abadi, T., Hudson, A., Zervos, A., Powrie, W. (2015). The use of under
sleeper pads to improve the performance of ballasted railway track at switches and
[77] Plotkin, D., Davis, D. Bridge approaches and track stiffness. Final Report,
[78] Priest, J. A., Powrie, W., Yang, L., Grabe, P. J., Clayton, C. R. I. Measurements of
60(9): 667-677.
182
[79] Puppala, A. J., Saride, S., Archeewa, E., Hoyos, L. R., Nazarian, S.
[80] Read, D., Li, D. Design of track transitions. Transit cooperative research program:
[81] Ribeiro, Alves. C., Paixão, A., Fortunato, E., & Calçada, R. (2015). Under sleeper
1449.
Exposition; 2006.
[83] Sasaoka, C. D., Davis, D. Implementing track transition solutions for heavy axle
load service. In Proceedings, AREMA 2005 Annual Conference, Chicago, IL; 2005.
[84] Sasaoka, C., Davis, D., Koch, K., Reiff, R., & GeMeiner, W. (2005). Implementing
[85] Sañudo, R., Dell'Olio, L., Casado, J. A., Carrascal, I. A., Diego, S. Track transitions
[86] Sakurai, T. , Nakamura, T. , and Muramoto, K. Cyclic Loading Test with Full-Scale
Model for Pre-Stressed Ballast Track. Geotechnical Pavement Research in Japan II,
Final Report WG4. Railway Technical Research Institute, Tokyo, 2013, pp. 175–
183
178. Reproduced in Geotechnical Pavement Research in Japan II—Final Report by
[87] Selig, E. T., & Li, D. (1994). Track modulus: Its meaning and factors influencing
[88] Selig, E. T., & Waters, J. M. (1994). Track geotechnology and substructure
[89] Schneider, P., Bolmsvik, R., & Nielsen, J. C. (2011). In situ performance of a
ballasted railway track with under sleeper pads. Proceedings of the Institution of
Mechanical Engineers, Part F: Journal of Rail and Rapid Transit, 225(3), 299-309.
[90] Scullion, T., Briggs, R. C., Lytton, R. L. Using the multidepth deflectometer to
[91] Sharpe, P., Armitage, R. J., Heggie, W. G., & Rogers, A. (2002, July). Innovative
Engineering.
[92] Silva, É. A., Pokropski, D., You, R., & Kaewunruen, S. (2017). Comparison of
structural design methods for railway composites and plastic sleepers and
184
[93] Smith, M. E., Bengtsson, P. E., & Holm, G. (2006). Three-dimensional analyses of
237-242.
[94] Stark, T. D., Wilk, S. T., Thompson, H. B., & Sussmann, T. R. (2015, June). Effect
2015 conference.
[95] Sussmann, T., Selig, E. Evaluation of increased axle loading on northeast corridor
[97] Tanabe, M., Wakui, H., Matsumoto, N., Okuda, H., Sogabe, M., Komiya, S.
140(1): 705-710.
[98] Tutumluer, E., Stark, T. D., Mishra, D., Hyslip, J. P. Investigation and mitigation
[99] Tutumluer, E., Mishra, D., Boler, H., Hyslip, J.P., Thomas, M. Dynamic Responses
185
[100] Varandas, J. N., Hölscher, P., Silva, M. A. Dynamic behaviour of railway tracks on
[102] Vorster, D. , Gräbe, P. J. The effect of axle load on track and foundation resilient
deformation under heavy haul conditions. India: New Delhi, In 10th international
[103] Wang, H., Markine, V. L., Shevtsov, I. Y., Dollevoet, R. Analysis of the dynamic
V001T01A024-V001T01A024.
[104] Wang, H., Markine, V. Methodology for the comprehensive analysis of railway
[105] Wang, H., Markine, V. Dynamic behaviour of the track in transitions zones
considering the differential settlement. Journal of Sound and Vsibration 2019; 459:
114863.
[106] Warren, B.J. (2015). Field application of expanding rigid polyurethane stabilization
U.S.
Champaign).
186
[108] Witt, S. (2008). The influence of under sleeper pads on railway track dynamics.
Sweden.
[109] White, D., Sritharan, S., Suleiman, M., Mekkawy, M., & Chetlur, S. (2005).
Identification of the Best Practices for Design, Construction, and Repair of Bridge
Approaches. Final Report, Iowa DOT Project TR-481, Center for Transportation
109-115.
[111] Woodward, P. K., Kennedy, J., Medero, G. M., & Banimahd, M. (2012).
Mechanical Engineers, Part F: Journal of Rail and Rapid Transit, 226(3), 257-271.
[112] Yang, Y. B, Yau, J. D., Hsu, L. C. Vibration of simple beams due to trains moving
[113] Yang, C., Sun, H., Zhang, J. Dynamic responses of bridge-approach embankment
transition section of HSR. Journal of Central South University 2013; 20(10): 2839-
2839.
[114] Yean, S. W. and Yeong, B. Y. Steady-state response and riding comfort of trains
25(2): 251-265.
187
[115] Youcef, K., Sabiha, T., Mostaf, D., Ali, D. and Bachi, M.. Dynamic analysis of
train-bridge system and riding comfort of trains with rail irregularities. Journal of
[116] Yin, C., Wei, B. Numerical simulation of a bridge-subgrade transition zone due to
2013, 15(2).
[117] Yu, Z. W., & Mao, J. F. (2018). A stochastic dynamic model of train-track-bridge
[118] Zaman, M., Gopalasingam, & A., Laguros, J. (1991). Consolidation settlement of
[119] Zhai, W., Sun, X. A detailed model for investigating vertical interaction between
railway vehicle and track. Vehicle system dynamics 1994; 23(S1): 603-615.
[120] Zhai, W., Cai, Z. Dynamic interaction between a lumped mass vehicle and a
987-997.
[121] Zhai, W., He, Z., Song, X. Prediction of high-speed train induced ground vibration
[122] Zhai, W., Xia, H., Cai, C., Gao, M., Li, X., Guo, X., Wang, K. High-speed train–
188
[123] Ziyaeifar, M. Vibration control in train–bridge–track systems. Vehicle System
extension://oemmndcbldboiebfnladdacbdfmadadm/https://www.infrastructurerepo
rtcard.org/wp-content/uploads/2017/01/Rail-Final.pdf
189
APPENDIX A : FIELD GPR DATA AND LAYER PROFILES
Figure A.1 (a) GPR Scans, (b) Mid-Chord Offset (MCO) Data, and (c) Space Curve for
Track 2 at Madison Street Bridge Approach.
190
Figure A.2 (a) GPR Scans, (b) Mid-Chord Offset (MCO) Data, and (c) Space Curve for
Track 2 at Caldwell Street Bridge Approach.
191
A.2 Substructure Layer Profiles
Figure A.3 Substructure Layer Profile for Madison Street 12 ft. from the South Abutment
(Track 2).
Figure A.4 Substructure Layer Profile for Madison Street 60 ft. from the South Abutment
(Track 2).
192
Figure A.5 Substructure Layer Profile for Caldwell Street 80 ft. from the South Abutment
(Track 3; West End of Tie).
Figure A.6 Substructure Layer Profile for Caldwell Street 80 ft. from the South Abutment
(Track 3; East End of Tie).
193
APPENDIX B : FIELD MONITORED TRANSIENT DATA
Figure B.1 Madison Open Track -Transient Vertical Deformation with Train Passage.
Figure B.2 Madison Open Track -Transient Wheel Load on Crib with Train Passage.
194
Figure B.3 Madison Open Track -Transient Wheel Load on Tie with Train Passage.
Figure B.4 Madison Near Bridge -Transient Vertical Deformation with Train Passage.
195
Figure B.5 Madison Near Bridge -Transient Wheel Load on Crib with Train Passage.
Figure B.6 Madison Near Bridge -Transient Wheel Load on Tie with Train Passage.
196
3.5
Opentrack_CI_high
2.5
Nearbridge_CI_low
2 Opentrack_CI_low
Nearbridge
1.5 Opentrack
1
4/1/2012 10/18/2012 5/6/2013 11/22/2013 6/10/2014 12/27/2014 7/15/2015
Time
Figure B.7 Madison Street -Peak Vertical Deformation over Time (2012-2015).
197
100
90 Nearbridge_CI_high
Peak Wheel Load on Tie (kN)
Opentrack_CI_high
80
Nearbridge_CI_low
70 Opentrack_CI_low
Nearbridge
60
Opentrack
50
40
9/14/2011 10/18/2012 11/22/2013 12/27/2014 1/31/2016
Time
Figure B.8 Madison Street-Peak Wheel Load on Tie over Time (2012-2015)
Table B.2 Madison Street - Statistics of Peak Vertical Load Measured on Tie
(2012-2015)
198
160
140
130
Nearbridge
120
Opentrack
110 Nearbridge_CI_high
100 Opentrack_CI_high
90 Nearbridge_CI_low
80 Opentrack_CI_low
70
4/1/2012 10/18/2012 5/6/2013 11/22/2013 6/10/2014 12/27/2014 7/15/2015
Time
Figure B.9 Madison Street -Peak Wheel Load on Crib over Time (2012-2015)
Table B.3 Madison Street - Statistics of Peak Vertical Load Measured on Rail Crib
(2012-2015)
199
B.2 Caldwell Street Field Transient Data
Figure B.10 West of Caldwell Open Track - Transient Vertical Deformation with Train
Passage.
Figure B.11 West of Caldwell Open Track - Transient Wheel Load on Crib with Train
Passage.
200
Figure B.12 West of Caldwell Open Track - Transient Wheel Load on Tie with Train
Passage.
Figure B.13 East of Caldwell Open Track - Transient Vertical Deformation with Train
Passage.
201
Figure B.14 East of Caldwell Open Track - Transient Wheel Load on Crib with Train
Passage.
Figure B.15 East of Caldwell Open Track - Transient Wheel Load on Tie with Train
Passage
202
Aug-12_1
2.5
Aug-12_2
Peak Transient Deformation (mm)
2.3
East of Open Track
West of Open Track
2.1
East_CI_high
West_CI_high
1.9
Aug-12_1 Aug-12_2 East_CI_low
West_CI_low
1.7
1.5
4/1/2012 10/18/2012 5/6/2013 11/22/2013 6/10/2014 12/27/2014 7/15/2015
Time
Figure B.16 Caldwell Street - Peak Vertical Deformation over Time (2012-2015)
203
120
Aug-12_1
Peak Wheel Load on Tie (kN) Aug-12_2
110
100
East of Open Track
Aug-12_2 West of Open Track
90
East_CI_high
80 Aug-12_1 West_CI_high
East_CI_low
70 West_CI_low
60
4/1/2012 10/18/2012 5/6/2013 11/22/2013 6/10/2014 12/27/2014 7/15/2015
Time
Figure B.17 Caldwell Street -Peak Wheel Load on Tie over Time (2012-2015)
Table B.5 Caldwell Street - Statistics of Peak Vertical Load Measured on Tie
(2012-2015)
204
160
Aug-12_1
Aug-12_2
140
East of Open Track
130 West of Open Track
East_CI_high
Aug-12_2
120 West_CI_high
Aug-12_1
East_CI_low
110
West_CI_low
100
4/1/2012 10/18/2012 5/6/2013 11/22/2013 6/10/2014 12/27/2014 7/15/2015
Time
Figure B.18 Caldwell Street -Peak Wheel Load on Crib over Time (2012-2015)
Table B.6 Caldwell Street - Statistics of Peak Vertical Load Measured on Rail Crib
(2012-2015)
205
B.3 Field Monitored Settlement Data
(a)
(b)
Figure B.19 (a) Layer Settlements at Madison Street 12 ft. from South Abutment (b)
Layer Settlements at Madison Street 60 ft. from South Abutment.
206
(a)
(b)
Figure B.20 Layer Settlements at Caldwell Street 80 ft. from South Abutment (a) West
End of Tie (b) East End of Tie
207
APPENDIX C : FOURIER TRANSFORM BASED MODEL
Following the work by Kalker (1996) and Huang et al. (2009), Fourier Transform
technique can be utilized to solve for the equations of motion of the system. The Fourier
Transform was first performed from time to frequency domain and from spatial domain to
Upland Street bridge approach location is used with its measured transient
deformations as an example to compare and validate the predictions by the track transition
model. Please note that these are the preliminary results of the track transition model and
In this case study, the track model included 40 ties spaced at 609.6 mm on the
embankment side and 15 ties spaced at 609.6 mm on the bridge side. The wheel load was
taken as 140,000 𝑁 for Acela Express locomotive moving at a constant speed of 50 m/s
(110 mph), in accordance with the strain gauge measurements at the Upland Street bridge
approach site. Note that in the field instrumentation, MDDs were installed at approximately
60 ft. away from the abutment of the bridge to correspond to the open track location and
15 ft. away from the abutment of the bridge for the bridge approach location. For further
validation of the developed analytical model with the field data, the load was set to start
from tie No.4 to tie No. 50. The observation point on the embankment side was at tie No.
10, which is 30 ties away from the bridge abutment (approximately 60 ft.). Figure C.1
shows a sketch of the loading and response measurement locations in the case study. Note
208
Figure C.1 Sketch of Loading and Response Measurement Location.
Figure C.2 shows the vertical transient displacement in the model under one
constant moving load on the embankment side. It can be seen that the moving load occurs
on top of the observation point at around 0.07 sec. The rail oscillates when the moving load
is first applied, followed by an increasing trend as the load is approaching the observation
point, and finally quickly dissipating to almost zero due to the natural frequency of rail
itself. The maximum amount of downward deflection of rail predicted by track transition
model is approximately 1.7 mm. The maximum amount of upward deflection predicted by
track transition model is approximately 0.3 mm. The results obtained by Fourier Transform
The results from Fourier Transform Data are used as a reference case for checking
209
-4 Open Track Side - Upland August 2012
x 10
20
15
Total Vertical Displacement (m)
10
-5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (sec)
Figure C.2 Rail Deflection under One Moving Load at Open Track Location.
210
APPENDIX D : CODE FOR TRAIN-TRACK-BRIDGE MODEL
This section documents the MATLAB code developed for establishing the train-
tic
%% Input parameters
Num_veh = 8;
211
%% Track structure parameters
1.24e5 as Zhai
1.2e9 as original
5.9e4 as Zhai
Zhai
Zhai
Huang
tie(tt)=tie(tt-1)+tiespace;
end
212
%% Vehicle parameters
l2=7.8994;
l3=2.8448;
l2_c=15.1384;
l3_c=2.9972;
pos = zeros(1,Num_veh*4);
locomotive
%% Locomotive
213
Mc = 45000; % Mass of Car body in kg 77000,90000
Mbogie2 = 1200;
Mw2 = 1200;
Mw3 = 1200;
Mw4 = 1200;
g = 9.8; %gravity
Jbogie2 = 760;
Kc2 = 5.32e6;
Cc1 = 4.9e4;
Cc2 = 7e4;
R=300;
%% Passenger car
Mbogie2_c = 1200;
Mw2_c = 1200;
214
Mw3_c = 1200;
Mw4_c = 1200;
Jbogie2_c = 760;
Kc2_c = 5.32e6;
Cc1_c = 4.9e4;
Cc2_c = 7e4;
%% Observation Point
tob=floor(Num_tie_track/2);
xob=(tob-1)*tiespace;
%%%%%% Locomotive
%Mass matrix
M=[Mc 0 0 0 0 0 0 0 0 0;
0 Mbogie1 0 0 0 0 0 0 0 0;
0 0 Mbogie2 0 0 0 0 0 0 0;
0 0 0 Mw1 0 0 0 0 0 0;
0 0 0 0 Mw2 0 0 0 0 0;
0 0 0 0 0 Mw3 0 0 0 0;
215
0 0 0 0 0 0 Mw4 0 0 0;
0 0 0 0 0 0 0 Jc 0 0;
0 0 0 0 0 0 0 0 Jbogie1 0;
0 0 0 0 0 0 0 0 0 Jbogie2];
%Stiffness matrix
%Damping matrix
216
0 0 0 -Cc1*b Cc1*b 0 0 0 2*Cc1*b^2 0;
%Mass matrix
M_c=[Mc_c 0 0 0 0 0 0 0 0 0;
0 Mbogie1_c 0 0 0 0 0 0 0 0;
0 0 Mbogie2_c 0 0 0 0 0 0 0;
0 0 0 Mw1_c 0 0 0 0 0 0;
0 0 0 0 Mw2_c 0 0 0 0 0;
0 0 0 0 0 Mw3_c 0 0 0 0;
0 0 0 0 0 0 Mw4_c 0 0 0;
0 0 0 0 0 0 0 Jc_c 0 0;
0 0 0 0 0 0 0 0 Jbogie1_c 0;
0 0 0 0 0 0 0 0 0 Jbogie2_c];
%Stiffness matrix
217
%Damping matrix
%% vehicle matrix
M_v = zeros(Num_veh*10,Num_veh*10);
C_v = zeros(Num_veh*10,Num_veh*10);
K_v = zeros(Num_veh*10,Num_veh*10);
for mm=1:Num_veh
if mm == 1|| mm==Num_veh
M_v((mm-1)*10+1:mm*10,(mm-1)*10+1:mm*10)=M;
else
M_v((mm-1)*10+1:mm*10,(mm-1)*10+1:mm*10)=M_c;
end
end
for mm=1:Num_veh
if mm == 1|| mm==Num_veh
218
C_v((mm-1)*10+1:mm*10,(mm-1)*10+1:mm*10)=C;
else
C_v((mm-1)*10+1:mm*10,(mm-1)*10+1:mm*10)=C_c;
end
end
for mm=1:Num_veh
if mm == 1|| mm==Num_veh
K_v((mm-1)*10+1:mm*10,(mm-1)*10+1:mm*10)=K;
else
K_v((mm-1)*10+1:mm*10,(mm-1)*10+1:mm*10)=K_c;
end
end
M1=diag(p/2*L*ones(Num_mode,1));
C1=zeros(Num_mode,Num_mode);
K1=zeros(Num_mode,Num_mode);
C2=zeros(Num_mode,Num_tie);
K2=zeros(Num_mode,Num_tie);
for i=1:Num_mode
for k=1:Num_mode
for mm=1:Num_tie
end
219
end
for mm=1:Num_tie_track
end
K1(i,i) = K1(i,i)+0.5*EI*L*(pi*i/L)^4;
end
M2=diag(mt);
C3=transpose(C2);
K3=transpose(K2);
C4=diag(cp+[cb;c_deck]);
K4=diag(kp+[kb;k_deck]);
C5_track=-diag(cb);
K5_track=-diag(kb);
C5_bridge=zeros(Num_tie_bridge,Num_tie_track);
K5_bridge=zeros(Num_tie_bridge,Num_tie_track);
C5 = [C5_track;C5_bridge];
K5 = [K5_track;K5_bridge];
M3=diag(mb);
C6=diag(cb+cs);
C7=transpose(C5);
K7=transpose(K5);
K6main=diag(kb+ks+2*ts);
220
K6up = diag(-ts(2:Num_tie_track),1);
K6down = diag(-ts(1:Num_tie_track-1),-1);
K6=K6main+K6up+K6down;
Mr=[M1,zeros(Num_mode,Num_tie+Num_tie_track);...
zeros(Num_tie,Num_mode),M2,zeros(Num_tie,Num_tie_track);...
zeros(Num_tie_track,Num_mode+Num_tie),M3];
Cr=[C1,C2,zeros(Num_mode,Num_tie_track);...
C3,C4,C5;...
zeros(Num_tie_track,Num_mode),C7,C6];
Kr=[K1,K2,zeros(Num_mode,Num_tie_track);...
K3,K4,K5;...
zeros(Num_tie_track,Num_mode),K7,K6];
MMM = [M_v,zeros(Num_veh*10,Num_mode+Num_tie+Num_tie_track);...
zeros(Num_mode+Num_tie+Num_tie_track ,Num_veh*10),Mr];
CCC = [C_v,zeros(Num_veh*10,Num_mode+Num_tie+Num_tie_track);...
zeros(Num_mode+Num_tie+Num_tie_track ,Num_veh*10),Cr];
KKK = [K_v,zeros(Num_veh*10,Num_mode+Num_tie+Num_tie_track);...
zeros(Num_mode+Num_tie+Num_tie_track ,Num_veh*10),Kr];
for f = 1:Num_veh
F1((f-1)*10+8 : f*10)=0;
end
221
%% Newmark Beta Integration Method
Gamma=0.5;
Beta=0.25;
a0=1/(Beta* dt^2);
a1=Gamma/(Beta* dt);
a2=1/(Beta* dt);
a3=1/(2*Beta)-1;
a4=Gamma/Beta-1;
a5=dt *(Gamma/Beta-2)/2;
a6=dt *(1-Gamma);
a7=Gamma* dt;
nstep = floor(T/dt);
for n=1:nstep
if n == 1
t = t + dt
% Track structure
for k=1:Num_mode
222
for v = 1:Num_veh
KKK((v-1)*10+4,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-
KKK((v-1)*10+5,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+2))/L);
KKK((v-1)*10+6,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+3))/L);
KKK((v-1)*10+7,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+4))/L);
KKK(Num_veh*10+k,(v-1)*10+4)=-Kvr*sin(pi*k*(Velocity*t-pos((v-
KKK(Num_veh*10+k,(v-1)*10+5)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+2))/L);
KKK(Num_veh*10+k,(v-1)*10+6)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+3))/L);
KKK(Num_veh*10+k,(v-1)*10+7)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+4))/L);
end
end
for II = 1:Num_mode
for JJ = 1:Num_mode
for p = 1:Num_veh*4
KKK(Num_veh*10+II,Num_veh*10+JJ) =...
KKK(Num_veh*10+II,Num_veh*10+JJ)+...
Kvr*(sin(pi*II*(Velocity*t-pos(p))/L)*sin(pi*JJ*(Velocity*t-pos(p))/L));
end
end
end
F_bar(:,n+1)=F_F(:,n+1);
223
u_u(:,n+1)=K_bar\F_bar(:,n+1);
else
t = t + dt
KKK = [K_v,zeros(Num_veh*10,Num_mode+Num_tie+Num_tie_track);...
zeros(Num_mode+Num_tie+Num_tie_track ,Num_veh*10),Kr];
% Track structure
for k=1:Num_mode
for v = 1:Num_veh
KKK((v-1)*10+4,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-
KKK((v-1)*10+5,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+2))/L);
KKK((v-1)*10+6,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+3))/L);
KKK((v-1)*10+7,Num_veh*10+k)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+4))/L);
KKK(Num_veh*10+k,(v-1)*10+4)=-Kvr*sin(pi*k*(Velocity*t-pos((v-
KKK(Num_veh*10+k,(v-1)*10+5)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+2))/L);
KKK(Num_veh*10+k,(v-1)*10+6)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+3))/L);
KKK(Num_veh*10+k,(v-1)*10+7)=-Kvr*sin(pi*k*(Velocity*t-pos((v-1)*4+4))/L);
end
end
for II = 1:Num_mode
for JJ = 1:Num_mode
for p = 1:Num_veh*4
KKK(Num_veh*10+II,Num_veh*10+JJ) =...
224
KKK(Num_veh*10+II,Num_veh*10+JJ)+...
Kvr*(sin(pi*II*(Velocity*t-pos(p))/L)*sin(pi*JJ*(Velocity*t-pos(p))/L));
end
end
end
vector
CCC*(a1*u_u(:,n)+a4*v_v(:,n)+a5*a_a(:,n));
u_u(:,n+1)=K_bar\F_bar(:,n+1);
a_a(:,n+1)=a0*(u_u(:,n+1)-u_u(:,n))-a2*v_v(:,n)-a3*a_a(:,n);
v_v(:,n+1)=v_v(:,n)+a6*a_a(:,n)+a7*a_a(:,n+1);
end
end
%u_u = [vehicle,q,ut,ub]
tspan = (0:dt:T);
Ur = zeros (1,length(tspan));
225
Urdot = zeros (1,length(tspan));
for k = 1:Num_mode
Ur = (Ur + sin(pi*k*xob/L)*u_u(Num_veh*10+k,:));
end
Ucardot = v_v(1,:);
226
Uw1dot = v_v (4,:); % Wheelset 1 velocity in m/s
% Pwheel =
%% Calculate Force
WRF= WRF1(floor((pos(1)+rob)/Velocity/dt)+1);
end
%% rail displacement
% xlabel('Time (sec)','FontSize',12,'FontName','Arial');
% 'FontName','Arial');
% %xlim([0,5]);
227
% %ylim([-0.2,1.8]);
% %title('Vertical Displacement')
% hold on
% %tie displacement
% plot(tspan,Ut*10^3,'r')
% hold on
%%
% % ballast displacement
% plot(tspan,Ub*10^3,'k')
% set(legend,...
% 'FontSize',12,...
% 'FontName','Arial');
% % xlim([0,5])
% % ylim([-0.2,1.8])
%maxd=max(abs(Ur*10^3));
% figure (2)
% xlabel('Time (sec)','FontSize',12,'FontName','Arial');
% 'FontName','Arial');
% % xlim([0,5]);
% % ylim([-10,50])
% %maxf=max(abs(RTF))
228