You are on page 1of 16

Influence of Under Sleeper Pads on Ballast Behavior Under

Cyclic Loading: Experimental and Numerical Studies


Sinniah K. Navaratnarajah, Ph.D. 1; Buddhima Indraratna, Ph.D., F.ASCE 2; and Ngoc Trung Ngo, Ph.D. 3
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Railway industries are placing greater emphasis on implementing fast and heavy haul corridors for bulk freight and commuter
transport in order to deliver more efficient and cost-effective services. However, increasing dynamic stresses from the passage of trains
progressively degrades and fouls the primary load-bearing ballast layer, which inevitably leads to excessive settlement and instability, damage
to track elements, and more frequent maintenance. Ballasted tracks are subjected to even greater stresses and faster deterioration in sections
where a reduced ballast thickness is used (e.g., bridge decks) or at locations where heavier concrete sleepers are used instead of lightweight
timber sleepers. The inclusion of under sleeper pads (USPs) at the base of a concrete sleeper is one measure used to minimize dynamic
stresses and subsequent track deterioration. In this study, cyclic loads from fast and heavy haul trains were simulated using a large-scale
process simulation prismoidal triaxial apparatus (PSPTA) to investigate the performance of ballast improved by USPs. The laboratory results
indicate that the inclusion of an elastic element at the harder interface of the concrete sleeper–ballast reduces the stresses transmitted to the
ballast and the underlying layers and minimizes the amount of deformation and degradation of the ballast. A three-dimensional finite-element
model was used to predict the behavior of ballast, and the influence of USPs on the stress–strain responses of ballast generally agree with the
experimental findings. DOI: 10.1061/(ASCE)GT.1943-5606.0001954. © 2018 American Society of Civil Engineers.
Author keywords: Under sleeper pads; Railroad ballast; Cyclic load; Deformation; Degradation; Numerical model.

Introduction Rail tracks are subjected to faster deterioration in sections where


a reduced ballast thickness is used and at locations where heavier
Rail networks in many countries play a prominent role in the con- concrete sleepers (ties) are used in conjunction with increased axle
veyance of bulk freight and commuter transport. The Australian loads (e.g., iron ore networks in Australian mining hubs). From a
Bureau of Infrastructure, Transport, and Regional Economics structural viewpoint, the damping capacity of the track at these lo-
(BITRE) forecast to the year 2030 suggests that road and rail freight cations is inadequate (Li et al. 2015). In recent years, the use of soft
volumes will more than double, with domestic demand for manu- synthetic rubber elements in track foundations to restore track resil-
factured goods underpinning much of road freight’s future growth ience and alleviate track damage has become increasingly popular
and an expected expansion in mineral exports supporting the (Zhai et al. 2004; Martins et al. 2009; Marschnig and Veit 2011;
growth in rail freight (BITRE 2008). Therefore, conventional bal- Alves Costa et al. 2012; Nimbalkar et al. 2012; Indraratna et al.
lasted rail tracks must be improved in order to cope with increasing 2014a; Sol-Sanchez et al. 2015; Navaratnarajah and Indraratna
demand for commuter transport and heavy haul freight networks 2017). Among a wide range of mitigation techniques, one option is
for minerals and agricultural goods. However, cyclic and impact to modify the vertical stiffness of track by attaching resilient rubber
loads generated by fast-moving heavy haul cause breakage, high pads (known as under sleeper pads, or USPs) under sleepers, espe-
lateral spread, and settlement of ballast, leading to rail–sleeper (tie) cially under heavy-duty concrete sleepers (Loy 2008; Marschnig
misalignment and loss of load-bearing capacity, and hence, compro- and Veit 2011; Indraratna et al. 2012; Nimbalkar et al. 2012; Sol-
mised safety and frequent track maintenance. Therefore, improving Sanchez et al. 2015). In addition to reducing track stiffness, USPs
railroad stability to operate high-speed passenger and heavy haul increase the contact area between sleeper and ballast, which helps to
freight services has become a key challenge not only in Australia distribute the wheel load to more adjacent sleepers and leads to a
but also in Europe, America, and Asia. reduction in stress transmitted from the sleepers to the substructure
layer (Witt 2008; Indraratna et al. 2014a; Sol-Sanchez et al. 2014;
1
Senior Lecturer, Dept. of Civil Engineering, Faculty of Engineering,
Abadi et al. 2015). Sleeper pads also even out the differences in
Univ. of Peradeniya, Peradeniya 20400, Sri Lanka. Email: navask@eng stiffness along transition zones and turnouts (Loy 2009). When soft
.pdn.ac.lk USPs were installed under hard concrete sleepers, ballast degrada-
2
Distinguished Professor of Civil Engineering and Research Director, tion decreased, the intervals at which it was necessary to tamp the
Centre for Geomechanics and Railway Engineering, Univ. of Wollongong, ballast were extended by 2–2.5 times, and the service life of ballast
Wollongong City, NSW 2522, Australia; Director, ARC Training Centre for increased by at least 100%, resulting in enormous savings in main-
Advanced Technologies in Rail Track Infrastructure, Univ. of Wollongong, tenance expenditures (Lakuši et al. 2010; Marschnig and Veit 2011).
NSW 2522, Australia (corresponding author). Email: indra@uow.edu.au Ngo et al. (2017b) studied the use of geo-inclusions for improved
3
Research Fellow, Centre for Geomechanics and Railway Engineering, performance of ballasted track substructure in the laboratory; field
Univ. of Wollongong, Wollongong City, NSW 2522, Australia. Email:
observation and computational modeling showed that a substantial
trung@uow.edu.au
Note. This manuscript was submitted on November 7, 2017; approved reduction in shear strains and particle breakage could be gained by
on April 25, 2018; published online on July 14, 2018. Discussion period using geo-inclusions in the track substructure. Danesh et al. (2017)
open until December 14, 2018; separate discussions must be submitted for investigated the effect of ballast degradation on its shear strength
individual papers. This paper is part of the Journal of Geotechnical and behavior and indicated that by increasing the degradation, i.e., the
Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. breakage ratio, up to 21%, shear strength decreases up to 48%.

© ASCE 04018068-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


In addition, the dilation angle of degraded ballast assemblies de- from the quarry was cleaned with water to remove any dust and
creases both with an increase in breakage ratio and with an increase clay adhering to the aggregates, dried before screening through
in normal stress; this is in agreement with the study conducted by selected sieve sizes, and then mixed in the desired proportions
Indraratna et al. (2011a). Currently, there is a lack of a comprehen- to obtain the required particle size distribution (PSD) in accor-
sive assessment of the most relevant geotechnical characteristics and dance with current Australian practice (AS 2015). Fig. 1 shows
the associated response of ballast under cyclic loading when resil- the PSD and grain size characteristics of the ballast. The maximum
ient elements are used in track substructure. Therefore, this research dry density γ d;max and corresponding optimum moisture content
was carried out to study the behavior of ballast under repeated cyclic womc of the subballast were 2,115 kg=m3 and 10%, respectively.
loading and to verify the effectiveness of USPs in attenuating the Prior to testing, the raw granular subballast was sieved to meet
adverse effects of cyclic stress–strain behavior and the correspond- Australian standards (AS 2015). Fig. 1 shows its PSD and grain
ing ballast degradation. size characteristics. The USP used in this study (see inset in Fig. 1)
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Laboratory experiments were carried out to simulate a proto- was a 10-mm-thick elastoplastic synthetic material made from
type track substructure using a novel process simulation prismoidal polyurethane elastomers. The USP was glued tightly to the base
triaxial apparatus (PSPTA) to investigate the behavior of ballast (790 × 220 mm) of the concrete sleeper (at the sleeper–ballast in-
with and without USPs under a large number of cyclic loadings terface) when testing with the pad. Fig. 1 shows the mechanical
(N ¼ 500,000 cycles), simulating axle loads of 25 and 35 t at properties of the USP. The USP material is predominantly elastic,
frequencies of 15 and 20 Hz (representing train speeds of 110 and with some plastic (irrecoverable) deformations observed on the pads
145 km=h, respectively). Numerical simulations using the com- after several years of use in tracks. The USP material does not have
mercially available finite-element program ABAQUS (2012) in any significant viscous (creep) behavior in the initial stages of its
three dimensions were also carried out to model the behavior of life, but viscous effects may develop after many years of use. The
ballast. The large-scale laboratory findings were used to calibrate aspect of potential viscous behavior over the long term was not
and validate the numerical model. The findings of this study will within the scope of the current study.
assist railway engineers in more effectively designing and con-
structing new tracks or in upgrading existing tracks adopting USPs
for concrete sleepers. Large-Scale Test Apparatus
The large-scale process simulation prismoidal triaxial apparatus
Experimental Study at the University of Wollongong, Australia was used to carry out
the cyclic loading tests. By exploiting the double symmetry of a
straight track section (Australian standard gauge), the plan dimen-
Laboratory Test Materials sions of the PSPTA (600 × 800 mm) represented a simplified unit
The fresh ballast and subballast materials commonly used in cell in which the side slope of the track prism and the role of
New South Wales (NSW) rail tracks were used for laboratory shoulder ballast have not been considered. The depth of the PSPTA
testing. These materials were obtained from Bombo quarry near was 600 mm. Also, unlike traditional test rigs with a fixed rigid
Wollongong. The ballast aggregates used in this study were pre- boundary (steel or Perspex), the PSPTA with movable vertical
dominantly latite (volcanic) basalt, a common igneous rock that steel walls allowed lateral displacements to occur against the lat-
can be found along the south coast of NSW, Australia. These dark eral confining pressure applied at the boundaries of the unit cell
basalt aggregates are fine-grained, very dense, and have sharp an- (Navaratnarajah and Indraratna 2017). The use of a Teflon spray
gular corners when blasted and quarried. The fresh ballast obtained on the sides of the steel wall surface eliminated undue shear stresses

Fig. 1. Particle size distribution and a sample under sleeper pad.

© ASCE 04018068-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


on the boundary, ensuring that the load was fully transmitted to constant at 15 kPa by a hydraulic jack connected to the vertical
the material being tested rather than to the boundary. Therefore, walls in the transverse direction. The accompanying lateral move-
the influence of the boundary can be considered minimal. A stan- ment ε3 was measured during testing [Figs. 2(b and c)]. The lateral
dard concrete sleeper of 220 mm wide and 200 mm thick confining pressure provided by the shallow granular foundation of
(Kaewunruen 2007) was cut to 780 mm long (i.e., slightly less a track is relatively small; therefore, the constant homogenous con-
than the chamber length of 800 mm) to fit inside the test chamber. fining pressure of 15 kPa applied was well within the acceptable
Figs. 2(a and b) show a schematic representation of the PSPTA, the range or variation measured in typical Australian tracks. Typically,
rail–sleeper assembly, and the substructure layers and dimensions. for NSW ballast gradations, the initial stresses were kept constant at
Fig. 2(c) shows the PSPTA cubical chamber. In this study, a plane around 10–15 kPa in the transverse direction (parallel to the
strain condition was simulated by ensuring a near-zero lateral strain sleeper) and at about 15–30 kPa along the longitudinal direction,
in the longitudinal direction (the train passage direction, which was for which the lateral strains had to be kept as small as possible
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

perpendicular to the sleepers). For a long, straight section of track, (i.e., plane strain) as measured by Indraratna et al. (2011b). This
longitudinal strain ε2 is negligible; thus, the assumption of a plane confinement was also validated by field measurements from Single-
strain condition (i.e., ε2 ¼ 0) is often reasonable. While the longi- ton and Bulli tracks in NSW and was also in agreement with large-
tudinal side walls were kept restrained by a hydraulic jack con- scale laboratory tests conducted by Le Pen (2008). Under a typical
nected to the wall, the confining pressure σ2 exerted by the walls confining pressure of 15 kPa, the PSPTA can allow a maximum
in the longitudinal direction varied during the tests, and it was mea- lateral movement of 50 mm (i.e., 25 mm on each side). Unlike tra-
sured by the load cell attached in line with the hydraulic jack ditional test rigs with a fixed rigid (steel or Perspex) boundary, the
[Figs. 2(b and c)]. The lateral pressure σ3 provided by the weight PSPTA, with its movable vertical walls (i.e., facilitated by small
of the crib and shoulder ballast in a real rail track was simulated by rolling castors), allows lateral displacements to occur against the
the lateral vertical walls (parallel to the sleeper) and was held lateral confining pressure applied at the boundaries of the unit cell.

Fig. 2. (Color) PSPTA and the test setup: (a) front view; (b) plan view; and (c) triaxial cubical chamber of PSPTA.

© ASCE 04018068-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Although in the field the lateral confinement provided by ballast not directly under the sleeper) were painted yellow [Figs. 3(b–d)].
shoulders varies when ballast starts to displace under the axle load, The spray paint used in this study was a non-oil-based product that
in this experimental procedure the lateral pressure was kept con- provided a very thin coating (bright colors) on the particles; there-
stant at 15 kPa. fore, the surface texture of the ballast was not affected. Large-scale
direct shear tests (300 mm width × 300 mm length × 200 mm height)
were carried out on both original (i.e., not painted) ballast and
Experimental Setup and Test Procedures painted-ballast assemblies by Indraratna et al. (2017); the tests in-
dicated that the paint’s influence on interparticle friction was neg-
The bottom layer of the test specimen consisted of a 150-mm-thick ligible (i.e., indicated almost the same shear strength envelope). The
subballast [compacted to around 95% of γ d;max using a rubber rail–sleeper assembly was then placed at the center [Fig. 3(d)] and
padded vibratory hammer; see Fig. 3(a)] overlain by a 300-mm- the space around the sleeper was filled with white-color-coated
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

thick layer of ballast. Fresh ballast was placed as three 100 mm thick [Fig. 3(e)] 150-mm-thick compacted ballast (crib ballast). When
layers and compacted to a typical field density of 1,560 kg=m3 (ini- testing with USPs, the pads were glued tightly to the base of the
tial void ratio e0 ¼ 0.73, and specific gravity Gs ¼ 2.7) by using a concrete sleepers [Fig. 3(f)]. A total of 500,000 load cycles (N) were
rubber padded vibratory compactor to minimize any ballast break- applied for each test. The different types of instrumentation and
age during tamping [Fig. 3(b)]. The representative ballast aggre- methods of data acquisition used in this study and the details of the
gates directly under the sleeper in each of the three layers (bottom, cyclic loads applied are explained in subsequent sections. Due to
middle, and top layers) were color coded red, green, and blue to the height limitation of the PSPTA (600 mm) the effect of subgrade
differentiate them and better assess the degradation of ballast with was not considered in this study. The depth of the test chamber was
depth [Figs. 3(b–d)]. The aggregates below the sleeper (which were sufficient to accommodate the typical depth of ballast in Australian

Fig. 3. (Color) (a) Compacted subballast layer; (b) compaction of ballast using a rubber padded vibratory hammer; (c) pressure cell placed at the
middle of the ballast layer; (d) rail–sleeper assembly placed on top of the ballast layer and pressure cell placed at the sleeper–ballast interface; (e) top
view of final test arrangements showing crib ballast around sleeper, settlement plates, potentiometer, and servohydraulic actuator; and (f) USP
attached to concrete sleeper.

© ASCE 04018068-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


track conditions (i.e., 250–300 mm) followed by the underlying sub- cyclic loading applied was determined using the AREA method (Li
ballast. No laboratory equipment is able to represent the actual depth and Selig 1998a) and is explained elsewhere (Navaratnarajah and
of the subgrade soil or rock below the track, and this limitation is Indraratna 2017). Initially, a strain-controlled loading at a rate of
also stated in the revised paper, where the nondisplacement boun- 1 mm=s was applied to bring the ballast–sleeper interface stress
dary of the PSPTA was fixed at a depth of 600 mm. Given that the 0
up to the mean cyclic deviator stress σ1cyc;mean of 130 and 175 kPa
bottom steel boundary was rigid and sitting on a strong laboratory for 25 and 35 t, respectively, and then a stress-controlled loading
0
floor (steel and concrete), this equipment was sufficient to mimic was applied by selecting loading amplitude A of ½1=2ðσ1cyc;max −
to some extent very stiff subgrade, but in situations where deep 0
σ1cyc;min Þ 100 and 145 kPa, corresponding to axle loads of 25 and
soft clays exist as subgrade, measured settlements will be underes-
35 t, respectively. The stress-controlled load was applied in two
timated by PSPTA as explained elsewhere by Indraratna et al.
stages: (1) a conditioning phase in which a reduced loading fre-
(2011b).
quency (f) of 5 Hz was used for 100 load cycles N to improve
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

contact at the sleeper–ballast interface and avoid losing actuator


Instrumentation and Data Acquisition contact with the rail top surface, which ensures that the different
components are seated properly at their interfaces; and (2) a loading
The vertical and lateral loads were measured using in-built load phase in which an actual test frequency (15 and 20 Hz) was applied
cells attached to the main actuator and to the hydraulic jacks, re- for the remaining 500,000 cycles.
spectively [Fig. 2(c)]. Lateral deformation was measured by the to-
tal lateral displacements of the two side walls. Linear variable
differential transformers were connected between the vertical walls Experimental Results and Discussion
and the fixed external frames to measure the lateral movements of
the walls. A system of hinges and linear ball bearings that con-
Sleeper–Ballast Interface Contacts
nected to the two side walls was regularly lubricated to minimize
frictional resistance and enable the walls to displace laterally with The performance of USPs in the attenuation of stress–strain and
minimum friction. Rapid response hydraulic pressure cells (with a degradation responses was evaluated by eight large-scale cyclic
capacity of 1 MPa, a diameter of 230 mm, and a height of 25 mm) load tests. When a track transmits a more concentrated rail seat load
specifically made for granular soil types were placed at (1) the bal- to sleepers, the sleeper–ballast interface distributes the load over
last–subballast interface [Fig. 3(a)], (2) the middle of the ballast a larger area. However, the sleeper only transmits stresses through
layer [Fig. 3(c)], and (3) the sleeper–ballast interface [Fig. 3(d)] a finite number of small, discrete contacts at the ballast–sleeper
in order to measure variations in pressure with depth. The pressure interface (Le Pen 2008). In this study, to increase the interface
cells were placed with care to ensure they were in full contact with contact area, a resilient rubber pad (USP) was used with the con-
the subballast, the ballast layers, and the concrete sleeper to avoid crete sleeper. Visual observations showed that the ballast particles
any erroneous data being recorded during testing. The settlement bedded uniformly with the concrete sleeper when a USP was placed
plates (base of 100 × 100 × 6 mm attached to 10-mm-diameter on the base of the sleeper. Fig. 4(a) shows the USP attached to the
rods with lengths that matched the depth of the embedded layers) base of the concrete sleeper before the test, and Fig. 4(b) shows the
installed at the ballast–subballast interface and the potentiometers
(stroke 100 mm) placed at the top of the sleeper were used to
measure the vertical deformation of the ballast [Figs. 3(d and e)].
Every instrument was calibrated and the calibration factors were
inserted into the host computer, which is supported by LabVIEW
(version 10.0) software to record all associated load, stress, and
deformation data against time and the number of loading cycles
N during testing. The transducer casings and the cables of the in-
struments were encased in flexible rubber conduits to protect them
from damage caused by vibratory compaction during preparation
and during the cyclic loadings. All sensors (pressure plates and po-
tentiometers) were connected to an electronic data logger DT800
(HBM, West Leederville, Australia) and were controlled by a host
computer supported by LabVIEW software to record continuous
data at a frequency of every 1 s. A total of half a million load cycles
was applied in each test, but the tests were interrupted at specific
cycles (100, 500, 1,000, 5,000, 10,000, 50,000, 100,000, 200,000,
300,000, 400,000, and 500,000) to take manual readings from set-
tlement pegs and to capture the resilience of ballast material at the
end of these cycles. The initial readings of all the instruments were
taken before starting the cyclic loading test phase. The final weights
of each of the three ballast layers (after screening through selected
sieves as per the PSD) were carefully measured for ballast breakage
calculations.

Applied Cyclic Loading


The cyclic load was applied to the rail–sleeper assembly and ballast
Fig. 4. (Color) USP (a) before test; and (b) after test (at the end of
layer by a servohydraulic actuator through a 100 mm diameter
500,000 load cycles).
cylindrical steel ram [Figs. 2(c) and 3(e)]. The harmonic sinusoidal

© ASCE 04018068-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


imprints of ballast on the pad after the test (i.e., after 500,000 load shown in Figs. 5(a and b), respectively. The residual plastic defor-
cycles). This clearly indicates that the contact area between the bal- mation of ballast was measured at the end of each selected cycle N
last and the sleeper was significantly increased, because the resil- by taking the average value of the difference between the readings
ient layer allowed the ballast aggregates to be partially pressed from the potentiometers placed on top of the sleepers and readings
(indented) at the surface of USP. This resulted in a more uniform from the settlement plates placed at the ballast–subballast interface,
distribution of pressure than would have occurred without a USP on to exclude any plastic deformation of the subballast layer.
the base of the concrete sleeper. Although polymeric materials such The test results indicated that the ballast deformed rapidly up
as sleeper pads deteriorate when exposed to the environment (from to around 10,000 cycles due to its initial densification and further
thermal exposure, oxidation, photolysis, and hydrolysis), they have packing after the corners of the sharp angular aggregates began to
a service life of approximately 20 years (Ito and Nagai 2008; break. However, once the ballast started to stabilize, the rate of de-
Lakuši et al. 2010). A careful visual check of the sleeper pads after formation gradually decreased and then remained relatively constant
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

the tests showed that the pads still remained in contact with the after 100,000 cycles. This showed that ballast aggregates underwent
sleeper and were in good condition [Fig. 4(b)] other than some plas- considerable particle rearrangement and densification during the in-
tic indentation marks on the surface of the pads. With trains passing itial cycles, but after reaching a threshold compression, resisted fur-
repeatedly, it is anticipated that the elastic behavior of sleeper pads ther deformation during any subsequent load cycles (Indraratna
may deteriorate to some extent over time (after several years); the et al. 2013). It is evident that ballast deformation (vertical and lat-
long-term viscous considerations of sleeper pads have not been eral) decreased significantly by using USPs. The increased contact
studied, and this is a limitation of this study. area of ballast by the USPs at the sleeper–ballast interface reduced
the induced stresses at the interface and at particle–particle contact,
allowing for a more uniform distribution of stress, which in turn
Vertical and Lateral Plastic Deformation and Strain decreased overall deformation. For the limited loading frequencies
(15 and 20 Hz) tested in this study, the vertical strain decreased in
The stress–strain response of ballast under cyclic loading with and
the range of 19%–29% for a 25-t axle load and about 21% for a 35-t
without resilient rubber mats was evaluated in this study. The ac-
axle load when USPs were used with the concrete sleeper. Lateral
cumulated vertical Svp and lateral Shp plastic deformations and cor-
strain decreased in the range of 9%–14% for a 25-t axle load and
responding vertical (ε1 ¼ Svp =300) and lateral (ε3 ¼ Shp =800)
9%–11% for a 35-t axle load, as shown in Figs. 5(a and b). This
plastic strains of ballast (with and without USP) measured at axle
means that attaching USPs at the base of concrete sleepers—for
loads of 25 and 35 t at loading frequencies of 15 and 20 Hz are
either new tracks or when wooden sleepers on an existing track
are being replaced by concrete sleepers—is an ideal solution for
reducing the long-term plastic settlement of a track subjected to
repetitive cyclic loading.

Influence of Axle Loads and Train Speed on Vertical


and Lateral Deformation of Ballast
The influence of axle loads and train speeds (loading frequency) on
the permanent vertical and lateral deformation of ballast can best be
presented by comparing those values at the end of testing (i.e., after
500,000 cycles). Figs. 6(a and b) show the variation of vertical and
lateral strains with train speed for 25- and 35-t axle loads with and
without USP. The results show that increasing train speed and axle
load significantly increased the vertical and lateral strains of ballast
and that the USPs played a major role in reducing these strains.
When train speed increased from 110 to 145 km=h, the resulting
vertical and lateral strains increased in the order of 51%–57% and
27%–33%, respectively, and when the axle load increased from 25
to 35 t, the resulting increases in vertical and lateral strains were in
the order of 37%–43% and 53%–60%, respectively, as shown in
Figs. 6(a and b) for the tests without USP. These comparisons
verify that train speed has more influence on vertical strains than
on lateral strains, because increasing train speed increases the
acceleration and vibratory compaction of ballast aggregates. How-
ever, axle load has more influence on lateral strain than on vertical
strain, because compacted ballast has an adequate bearing capacity
for higher loads but, because of lower confinement, the extent of
lateral spreading is comparatively greater. Nevertheless, vertical
and lateral strains decreased markedly when USPs were placed
under the concrete sleepers.

Influence of Loading Frequency f and


Axle Load on Ballast Breakage
Fig. 5. (Color) Variation of (a) vertical; and (b) lateral plastic deforma-
Progressive particle degradation occurs when a repetitive cyclic
tion (and strain) with number of cycles N with and without USP.
load is applied to the ballast mass. Initially, grinding and corner

© ASCE 04018068-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


this was followed by the middle layer and then the bottom layer.
The BBI values obtained in this study were similar to those for
samples collected from the field prior to track maintenance: BBI
in the Bulli test track section (Ngo 2012) was 0.08–0.11, and
BBI in the Singleton track section (Indraratna et al. 2014b) was
0.06–0.17 for train speeds of about 70–80 km=h (approximately
10 Hz). The results of the present study indicate that ballast deg-
radation was significantly reduced by the USPs used under the con-
crete sleepers. On average, when all three layers are considered,
USPs reduced degradation by more than 50%. Considering each
ballast layer separately, the reduction for the 25- and 35-t axle loads
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

tested at 15 and 20 Hz frequency was in the order of 53%–63% for


the top layer, 51%–59% for the middle layer, and 48%–62% for the
bottom layer. The increased contact area due to USPs and the sub-
sequent reduction in the interface and interparticle ballast stress
is probably the main reason behind this substantial reduction in
ballast breakage.
Like plastic strain, ballast degradation also increases with train
speed (loading frequency) and axle loads. For the same axle loads
(25 or 35 t), increasing loading frequency from 15 to 20 Hz caused
an increase in the average BBI of all three layers in the order of
26%–28%. For the same loading frequency (f ¼ 15 or 20 Hz), in-
creasing axle loads from 25 to 35 t (such as from a passenger train
to a freight or coal train) increased the average BBI in the order of
34%–37%. Unlike vertical plastic strain, axle load has more influ-
ence on ballast breakage than loading frequency (i.e., train speed),
because a heavier load contributes more to ballast grinding, corner
breakage, and particle splitting. It can, therefore, be concluded that
increased train speed contributes more to plastic strain, whereas
increased axle load contributes more to ballast degradation.

Fig. 6. (Color) Variation of (a) vertical; and (b) lateral plastic deforma- Influence of USP on the Stress Distribution
tion (and strain) with train speeds for 25- and 35-t axle loads.
The pressure cells placed at the sleeper–ballast interface, in the
middle of the ballast layer, and at the ballast–subballast interface
were used to measure the dynamically induced pressure variations
breakage of angular ballast at the sleeper–ballast and interparticle along the depth of the ballast layer. Fig. 8 shows the variations in
contacts take place, followed by complete fracture across the body pressure with and without USPs. The load at the wheel–rail inter-
of the particles, depending on the strength of the parent rock and the face produced by a moving train on a real track is greater than that
level of stress increment (Indraratna et al. 2005; Lackenby et al. of the same train in a stationary state (Esveld 2001; Kerr 2003).
2007). This breakage of ballast particles contributes to increased Consequently, the dynamic stresses induced at the sleeper–ballast
axial and lateral strains and also causes differential track settlement. interface and stresses in the substructure layers below the sleeper
In this study, the ballast breakage index (BBI) method proposed are also higher. In conventional railway engineering practice, the
specifically for railway ballast by Indraratna et al. (2005) was used dynamic amplification factor (DAF) is used to define the equivalent
to evaluate ballast degradation. Since ballast breakage potential dynamic stresses, in which DAF is a function of static wheel load
varies with depth due to changes in stress along the depth of the and train speed (Li and Selig 1998b; Esveld 2001). The amplified
ballast layer, ballast mass was divided into top, middle, and bottom stresses shown in Fig. 8 correspond to dynamic stresses measured
layers and analyzed separately. These layers correspond to ballast for static loads applied at frequencies of 15 and 20 Hz (train speeds
breakage governed by variations in stress and type of interface. of 110 and 145 km=h, respectively), which are much larger than the
The top layer is the sleeper–ballast or USP–ballast zone, the middle static stresses.
layer is the ballast–ballast zone, and the bottom layer is the ballast– It is evident that the dynamically induced stresses in the ballast
subballast zone. The fresh ballast aggregates directly under the increased substantially with axle load and more moderately with
sleepers were color-coded to better quantify the degradation of train speed (loading frequency). When USPs were placed under
ballast in the three different zones; after testing, each layer of de- the concrete sleepers, there was a definite reduction in stress within
graded ballast was separated according to color to evaluate its BBI. the ballast mass, because the dynamically induced stress was re-
Fig. 7(a) shows photographs of fresh and degraded ballast at the duced by the softer interface between the ballast and the concrete
top, middle, and the bottom layers. The BBI computed in this study sleepers by the USPs. On average, the percentage reduction of
was based on overall ballast breakage at the completion of the tests stress for a 25-t axle load varied from about 12% at the sleeper–
(i.e., at the end of 500,000 load cycles). The BBI for the top, ballast interface to 10% at the middle of the ballast layer and to
middle, and bottom layers are presented in Figs. 7(b and c) for about 9.5% at the ballast–subballast interface. For a 35 t axle load,
ballast tested with and without USP for 25- and 35-t axle loads, the reduction in stress was about 10% at the sleeper–ballast inter-
respectively, at f ¼ 15 and 20 Hz. face, about 8.5% at the middle of the ballast layer, and about 8% at
As expected, ballast breakage was greater at the top layer be- the ballast–subballast interface. This reduction in stress was higher
cause of higher interface and interparticle contact stress at the top; in the zone near the USP–ballast interface. When a USP is used

© ASCE 04018068-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. (Color) (a) Color-coded fresh and degraded ballast from the top, middle, and bottom layers; and variation of ballast breakage index with
loading frequency and with and without USP for (b) 25 t; and (c) 35 t axle loads.

Fig. 8. (Color) Variation of ballast pressure with depth with and without USP.

with a concrete sleeper, the ballast aggregates make a more uniform interface (ballast–sleeper) contact when USPs were provided at
and increased area of contact at the sleeper–ballast interface, the base of sleepers leads to a reduction in interface stresses.
which leads to a reduction in ballast stress. These results are in The vertical stress distribution obtained from the conventional elas-
agreement with the findings by Dahlberg (2010) and Le et al. tic analysis (Boussinesq’s method) has been included in Fig. 8 for
(1999), in which the authors found that an increased area of comparison. These data correspond to 25- and 35-t axle loads at a

© ASCE 04018068-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


frequency of 20 Hz under without USP conditions. The theoretical summarizes the material properties used for the FE simulation.
elastic stress distribution obtained from Boussinesq’s method devi- There are two forms of bulk viscosity (linear and quadratic bulk
ates substantially from what was measured. It is noteworthy that a viscosity) introduced in the ABAQUS/Explicit analysis to improve
significant amount of energy dissipates through particle breakage, a the modeling of high-speed dynamic events. Therefore, a small
fact that the conventional elastic analysis cannot capture. amount of bulk viscosity was included by default in the FE analysis,
which is appropriate for most models (ABAQUS 2012). Surface-to-
surface contact elements were used to model the contact between
Numerical Analysis two surfaces.
The cyclic loading was applied at the rail seat to simulate
The stress–strain behavior of an integrated layered system con- 25- and 35-t axle loads with f ¼ 15 Hz. The mean [σ1cyc;mean ¼
sisting of a sleeper (tie), ballast, subballast, and under sleeper pad 1 1
(USP) has been investigated using the finite-element method, for 2 ðσ1cyc;max þ σ1cyc;min Þ] and amplitude [A ¼ 2 ðσ1cyc;max − σ1cyc;min Þ]
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

of the applied cyclic load was based on cyclic stress measured


which the commercially available software ABAQUS (2012) was
through data-burst techniques at regular intervals during actual
used. The three-dimensional (3D) FEM model developed in this
laboratory testing. A periodic type of amplitude curve with a sinus-
study simulates ballast behavior with and without a USP attached
oidal variation of stress with time was used under the ABAQUS
to the base of a concrete sleeper subjected to cyclic loading. The
load module to apply cyclic stress on the rail seat surface. The
model represents a plane strain condition by ensuring zero lateral
FE analyses were carried out only up to 10,000 load cycles for each
strain in the longitudinal direction (direction of train passage) as
set of simulations in consideration of the time required for solving a
explained previously. In this study, the dynamic explicit numerical
3D finite element problem under a large number of cyclic loadings
technique was used to simulate the repeated cyclic loading behavior
(Indraratna and Nimbalkar 2013).
of ballast.

Model Geometry and Boundary Conditions Finite-Element Discretization and Element Type
In the finite-element (FE) analysis, the unit cell of a rail track sub- The finite-element model was meshed using a structured pattern
structure (prototype sample) replicated in the PSPTA used in the consisting of 3D, 8-noded linear brick, and reduced integration
laboratory was modeled. Fig. 9(a) shows a full-size model of the (C3D8R) hexahedral elements. The computation time for a coarser
unit cell and its dimensions. To optimize computational time, only mesh is quite small but generally underestimates the displacements,
one-quarter of the unit cell was modeled in the FEM analysis by whereas a finer mesh is computationally more time-consuming and
exploiting the symmetry of the unit cell in each perpendicular di- expensive but provides more realistic results. Thus, a sensitivity
rection in the horizontal plane, as shown in Fig. 9(b). The boundary analysis to compare the vertical deformation obtained from differ-
conditions were applied by considering the geometric symmetry. ent numbers of elements of the same model was carried out by sim-
The interior vertical boundaries on either side of the axes of sym- ulating a 25-t axle load at f ¼ 10 Hz. The cyclic load of up to
metry were restrained from lateral movement in their respective di- 1,000 load cycles was applied with USPs for different numbers
rections and were only allowed to move in the vertical direction. of elements in the mesh ranging from 672 to 79,200. Based on
The bottom boundary was fixed for any movement in the vertical the results of this sensitivity analysis, an FE model with 10,200
direction to model a rigid boundary underlying the integrated sub- C3D8R elements (11,760 nodes) was selected for further cyclic
structure layers. A plane strain condition was applied to the exterior loading simulation. Fig. 9(c) shows the selected mesh with differ-
boundary in the longitudinal direction (intermediate principal ent types of substructure layers. The vertical size of the USP ele-
strain, ε2 ¼ 0), and the exterior transverse boundary was allowed ments were seeded locally to match the thickness of the USPs
to move with a lateral confining stress of 15 kPa (minor principal (i.e., 10 mm); all the other elements were >10 mm or mostly equal
stress, σ3 ) applied on the right boundaries of the exterior elements to 20-mm-thick cube elements.
on the right side. The top boundary of the model was allowed to
move in the vertical direction (free boundary) under the application
of cyclic stress (major principal stress, σ1cyc ) at the rail seat surface Numerical Analysis: Results and Discussion
[Fig. 9(b)]. In this numerical analysis, 25- and 35-t axle loads with a
loading frequency of 15 Hz were simulated with and without USP. Vertical Deformation
Figs. 10(a and b) show the vertical deformation Sv of a 35 t axle
Material Properties and Cyclic Loading
load simulated at a cyclic loading frequency f of 15 Hz up to
The ballast and subballast were modeled as an elastoplastic material 10,000 load cycles without and with USP, respectively. These fig-
with nonassociative behavior obeying the Drucker-Prager yield ures represent the deformation contours of the full model and a
criterion (Drucker and Prager 1952). This yield criterion is com- longitudinal section (a half-section along the middle of the sleeper)
monly used to simulate granular materials such as ballast, because in order to better understand the deformation behavior of the sub-
its strength and yield characteristics depend on the level of stress and structure layers. The same Sv stress limits for the contour plots are
the volumetric strain (Leshchinsky and Ling 2013; Biabani et al. used for Figs. 10(a and b) in order to compare deformation under
2016). The Drucker-Prager yield criterion, with a rounded yield sur- two different substructure conditions (i.e., with and without USPs).
face, is simple to use as opposed to the Mohr-Coulomb yield cri- These figures show that maximum Sv occurs in the vicinity of the
terion, with a sharp, hexagonally shaped yield surface around the loaded area of a sleeper. The ballast at the corners and longitudinal
hydrostatic axis (Leshchinsky and Ling 2013). The material param- edges of the unit cell showed a dilation behavior, because the crib
eters for the Drucker-Prager model were obtained based on labora- ballast was not confined from the top and was free to move in a
tory triaxial testing carried out on ballast and subballast materials vertical direction (i.e., free boundary at the top). The magnitude of
obtained from the Bombo quarry, New South Wales (Indraratna Sv decreased along the depth of the substructure layers. These re-
et al. 2011b). For simplicity, the concrete sleepers and the under sults also confirm that USPs placed under concrete sleepers sub-
sleeper pads were modeled as linear elastic materials. Table 1 stantially reduce vertical deformation of the substructure.

© ASCE 04018068-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. (Color) (a) Unit cell of a rail track substructure; (b) finite-element idealization of a quarter unit cell model showing loads and boundary
conditions; and (c) FE mesh for simulation and type of element.

To compare the vertical deformation Sv of ballast predicted by (with USP). The magnitude of Sv increased with the axle load, with
FE analysis with the laboratory data, the variation of Sv was plotted or without USP. The predicted value of Sv compares reasonably
against the number of cycles N for 25- and 35-t axle loads at well with the experimental results, at least at the initial stages of
f ¼ 15 Hz, as shown in Fig. 11(a) (without USP) and Fig. 11(b) loading cycles (N ≤ 1,000). Cyclic strain hardening appears in

© ASCE 04018068-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Table 1. Material properties for FE analysis Sh with and without USP, the same stress limits were used in these
Material stress contour plots. The results from the FE analysis confirm that
the use of USPs under concrete sleepers decreases the lateral
Under
spreading of the ballast and subballast layers.
Concrete sleeper
Property Ballast Subballast sleeper pad As Figs. 12(a and b) show, the magnitude of Sh increased with
3
depth, and the maximum Sh occurred at a depth of 140–160 mm
Density, γ (kg=m ) 1,560 2,115 2,400 420 from the base of the sleeper and then decreased in the deeper part
Elastic modulus, E (MPa) 125 100 36,000 6
of the substructure. The maximum lateral displacement contours
Poisson’s ratio, ν 0.25 0.3 0.2 0.48
Friction angle, ϕ (degrees) 45 38 — — concentrated in the vicinity of one-half to two-thirds depth from
Angle of dilation, ψ (degrees) 15 9 — — the base of the sleeper. These results suggest that to reduce lateral
spreading, the optimum location for any geo-inclusions (geogrids
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

or geocells) within the ballast layer would be at a depth between


one-half and two-thirds from the base of the sleeper. These obser-
the later stages of the loading cycles (N > 1,000); the predicted vations agree with the findings of Oh (2013) and Ferreira and
deformation curve flattens after about 1,000 cycles. The degrada- Indraratna (2017).
tion of ballast particles was not considered in this FE simulation, The average lateral deformations Sh of the transverse boundary
and this, in addition to the simplifications and approximations of the FE model and laboratory experiments were compared by
adopted in the numerical procedures, could be one of the reasons plotting the variations of Sh against the number of cycles for 25-
for the difference in deformation between the experimental and and 35-t axle loads at f ¼ 15 Hz. Figs. 13(a and b) show the results
numerical results at N > 1,000, (Leshchinsky and Ling 2013; without and with USPs, respectively. The predicted Sh by FE analy-
Biabani 2015; Sun 2015). Nevertheless, the magnitude of Sv sis confirms that increasing the axle load from 25 to 35 t increases
was markedly reduced by USPs for 25- and 35-t axle loads simu- the lateral spreading of ballast with or without the use of USPs,
lated at f ¼ 15 Hz, as captured well in this FEM simulation. which is consistent with the laboratory findings. The value of Sh
predicted by FEM was comparable to the experimental findings
despite the fact that in the FE analysis, most of the lateral defor-
Lateral Deformation mation occurred at the initial stages of the cyclic load (N ≤ 1,000);
Figs. 12(a and b) show the lateral deformation Sh of a 35 t axle load the laboratory results indicated an incremental lateral deformation
simulated at f ¼ 15 Hz up to 10,000 load cycles without and with even after N > 1,000. As previously mentioned, the fact that ballast
USPs, respectively. For the purpose of comparing the variation of breakage was not considered in the FE simulation could be one of

Fig. 10. (Color) Vertical deformation predicted from FEM (35 t axle load and f ¼ 15 Hz): (a) without USP; and (b) with USP.

© ASCE 04018068-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. (Color) Variation of vertical deformation of ballast with N: laboratory measurements versus FE model prediction (a) without USP; and
(b) with USP.

Fig. 12. (Color) Lateral deformation predicted from FEM (35 t axle load and f ¼ 15 Hz): (a) without USP; and (b) with USP.

the reasons for this. Nevertheless, the reduction of Sh was evident at f ¼ 15 Hz (1) for the full ballast layer, (2) for a longitudinal
from the simulation results when a USP was used with a concrete section (along the center of the sleeper), and (3) for a transverse
sleeper for 25- and 35-t axle loads simulated at f ¼ 15 Hz. section (across the center of the sleeper). These contour plots cor-
respond to the magnitude of the applied cyclic stress σ1cyc as it
reaches a maximum value σ1cyc;max . The stress contours indicate
Distribution of Stress in the Ballast
that the sleeper–ballast interface stress concentrates along the edges
Fig. 14 shows the distribution of vertical stress in ballast, with and of the sleeper, which is consistent with the contact force distribu-
without USP, obtained from the FE analysis for 35 t axle simulated tions observed by Ngo et al. (2017a). As expected, the stresses were

© ASCE 04018068-12 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. (Color) Variation of average lateral deformation of ballast with N, laboratory measurements versus FE model prediction: (a) without USP;
and (b) with USP.

higher in ballast directly under the rail seat, where the wheel load is matched reasonably well with the laboratory data, but the stress
applied. The average stress in the ballast decreased with the depth at the bottom (ballast–subballast interface) was overpredicted by
of the ballast layer. the FEM analysis. Ballast breakage during laboratory testing sig-
The contours of stress distribution presented in Figs. 14(a–c) nificantly influenced the stress measured at the bottom, because,
were captured at the end of the analysis (N ¼ 10,000). The ob- when the corners and edges of sharp and angular ballast particles
served stress was slightly higher at the initial cycles up to about broke into smaller pieces due to cyclic stress, they gradually moved
N ¼ 500 and then stabilized and remained almost constant for the toward the bottom of the ballast and filled the voids (i.e., vibratory
remaining load cycles. Nevertheless, the distribution of vertical compaction) and increased the interparticle contacts at the bottom
stress with and without USP in Figs. 14(a–c) clearly shows the ben- layer. This increase in contact area reduced the stress at the bottom
efits of using USPs with concrete sleepers. When USPs are used, layer, as was evident from the experimental results. However, since
the highly concentrated stresses at the sleeper–ballast interface ballast breakage was not considered in the current FE analysis, the
(from a hard concrete sleeper) are distributed over a wider area, stress predicted at the bottom of the ballast did not correctly match
which reduces the magnitude of the vertical stress at the sleeper– with the experimental findings. Moreover, the type of element
ballast interface and within the ballast mass at particle–particle (C3D8R) used to build the FEM mesh was a continuum element,
contacts. The results indicate that a USP with a concrete sleeper whereas rail ballast consists of a skeleton of aggregates, which dis-
substantially reduces the stresses transmitted to the ballast. The ver- tributes the stress via a discrete number of interparticle contacts
tical stress at the edges of the sleeper decreases along the depth of to larger areas toward the deeper regions of the ballast layer. Never-
ballast; the stress along the center line of the ballast layer increases theless, the decrease in stress in the ballast when a USP is used with
to a point at which the concentrated stress contours at the sleeper a concrete sleeper was correctly captured by the current FE analy-
edges reach the center line of the ballast mass and then decreases at sis, matching the experimental data.
greater depths [see transverse sections in Fig. 14(c)].
The maximum vertical stresses measured at the top (sleeper–
Conclusions
ballast interface), middle (ballast–ballast interface), and bottom
(ballast–subballast interface) of the ballast layer during laboratory The performance of ballasted track stabilized with under sleeper
testing were compared to the stress distribution from the FE analy- pads attached to the base of concrete sleepers was evaluated by
sis, as shown in Fig. 15(a) (without USP) and Fig. 15(b) (with experimental study and three-dimensional finite-element analysis.
USP). The variation of stress in the ballast along the center line A prototype track substructure consisting of rail, sleeper, ballast,
and along the edge of the sleeper was plotted as depicted by the and subballast layers was integrated inside the prismoidal chamber
cross-sectional view of the unit cell [see insets in Figs. 15(a and b)]. of the PSPTA and the performance of ballast was investigated with
As explained previously, the stress along the edge of the sleeper and without USPs placed under the concrete sleeper. Two axle
decreased from a maximum value at the sleeper–ballast interface, loads of 25 and 35 t with loading frequencies f of 15 and 20 Hz
while the stress along the center line peaks just above the middle of were tested under a large number of cyclic loading (N ¼ 500,000).
the ballast layer [Figs. 15(a and b)]. The “FEM-Average” shown The experimental findings showed promising outcomes in the
in Figs. 15(a and b) is based on the average values of 10 elements reduction of stress, vertical ε1 and lateral strain ε3 and degradation
(5 elements each along two center lines), which cover a total area of of ballast stabilized with USP. Vertical deformation decreased in the
200 mm diameter to approximately match the pressure cell cover- range of 19%–29% for the 25-t axle load, and about 21% for the
ing area of 230 mm diameter. The average ballast stress decreased 35 t axle load. The lateral deformation of the ballast decreased in
with the depth of the ballast layer. the range of 9%–14% and 9%–11% for 25- and 35-t axle loads,
The numerically predicted stress (FEM-Average) at the top respectively, when USPs were used with the concrete sleepers.
(sleeper–ballast interface) and middle (ballast–ballast interface) Axle loads and train speed (loading frequency) have a significant

© ASCE 04018068-13 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. (Color) Comparison of the distribution of vertical stress in the ballast layer with and without USP for a 35 t axle load simulated at f ¼ 15 Hz:
(a) full ballast layer; (b) longitudinal section; and (c) transverse section.

influence on ballast deformation. The results at N ¼ 500,000 indi- layer of ballast. Ballast degradation was reduced significantly when
cate that when the axle load increases from 25 to 35 t, vertical USPs were used at the hard concrete sleeper-to-ballast interface. On
and lateral deformations increase in the order of 37%–43% and average, when all three layers are considered, USPs reduced BBI by
53%–60%, respectively. When the train speed increases from 110 more than 50%. Considering each ballast layer separately, the reduc-
to 145 km=h (f ¼ 15 to 20 Hz), vertical and lateral deformations tion of BBI for the 25- and 35-t axle loads tested at f ¼ 15 and
increase in the order of 51%–57% and 27%–33%, respectively. 20 Hz were in the order of 53%–63% at the top layer, 51%–59%
Train speed (cyclic loading frequency) has more influence on ε1 at the middle layer, and 48%–62% at the bottom layer. The average
than on ε3 , whereas axle load has more influence on ε3 . BBI increased in the order of 26%–28% when f increased from 15
The USP leads to a significant increase in sleeper–ballast inter- to 20 Hz, whereas the percentage increase of BBI when axle load
face contacts and markedly reduces the stresses in the ballast, es- increased from 25 to 35 t was 34%–37%. Axle load had more in-
pecially at the top zone near the sleeper–USP–ballast interface fluence on ballast breakage than loading frequency (i.e., train
(10%–12%), followed by lesser reductions in the middle of the speed), because a heavier load contributes more to ballast grinding,
ballast layer (8.5%–10%) and the bottom zone near the ballast– corner breakage, and particle splitting.
subballast interface (8%–9.5%); these reductions lead to a reduction Although the particle breakage behavior of ballast was not in-
in ballast degradation. Regardless of the axle loads, frequencies, and corporated in the 3D FE analysis, the predicted vertical and lateral
USP condition, ballast breakage was greater at the top layer due to a deformation and the vertical stress distribution in the ballast layer
higher interface and interparticle contact stress at the top; there was generally agreed with the experimental findings—particularly the
less ballast breakage in the middle layer and even less in the bottom vertical and lateral deformation at N < 1,000 and the generated

© ASCE 04018068-14 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 15. (Color) Distribution of vertical stress in the ballast layer: laboratory measurements versus FE model prediction: (a) without USP; and
(b) with USP.

stresses at the top and middle of the ballast layer. The FE results References
indicated that USPs significantly reduce ε1 and ε3 and the vertical
stress of the ballast. Based on this study, using USPs on concrete Abadi, T., L. Le Pen, A. Zervos, and W. Powrie. 2015. “Measuring the area
sleepers is a favorable solution to reduce the stress–strain and deg- and number of ballast particle contacts at sleeper/ballast and ballast/
subgrade interfaces.” Int. J. Railway Technol. 4 (2): 45–72.
radation of ballast, and, in a practical sense, to minimize the main-
ABAQUS. 2012. ABAQUS documentation. Providence, RI: Dassault
tenance costs of ballasted rail tracks. Systèmes Simulia.
Alves Costa, P., R. Calçada, and A. Silva Cardoso. 2012. “Ballast mats for
the reduction of railway traffic vibrations. Numerical study.” Soil Dyn.
Acknowledgments Earthquake Eng. 42 (Nov): 137–150. https://doi.org/10.1016/j.soildyn
.2012.06.014.
The financial support provided by Australian Research Council AS (Standard Australia). 2015. Aggregates and rock for engineering
(ARC) and University Postgraduate Award (UPA) by the Univer- purposes, Part 7: Railway ballast. AS 2758.7. Sydney, Australia: AS.
sity of Wollongong to conduct this research study is gratefully Biabani, M. 2015 “Behaviour of geocell-reinforced subballast under cyclic
appreciated. Continuous assistance provided by Dr. Sanjay loading in plane strain condition.” Ph.D. thesis, Faculty of Engineering
and Information Sciences, Univ. of Wollongong.
Nimbalkar and A/Prof. Cholachat Rujikiatkamjorn during the study
Biabani, M. M., B. Indraratna, and N. T. Ngo. 2016. “Modelling of geocell-
is appreciated. The assistance provided by senior technical officers, reinforced subballast subjected to cyclic loading.” Geotext. Geomembr
namely Alan Grant, Cameron Neilson, and Ritchie McLean in 44 (4): 489–503. https://doi.org/10.1016/j.geotexmem.2016.02.001.
the School of Civil and Environmental Engineering, University BITRE (Bureau of Infrastructure, Transport, and Regional Economics).
of Wollongong, is also appreciated. 2008. Statistics yearbook 2007: Australian transport. Canberra

© ASCE 04018068-15 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068


ACT, Australia: Bureau of Infrastructure, Transport, and Regional Le, R., B. Ripke, and M. Zacher. 1999. “Ballast mats on high speed
Economics. bridges.” In Proc., 4th European Conf. on Structural Dynamics:
Dahlberg, T. 2010. “Railway track stiffness variations: Consequences and EURODYN’99, 699–703. Rotterdam, Netherlands: A.A. Balkema.
countermeasures.” Int. J. Civ. Eng. 8 (1): 1–12. Le Pen, L. 2008. “Track behaviour: The importance of the sleeper to ballast
Danesh, A., M. Palassi, and A. A. Mirghasemi. 2017. “Evaluating the interface.” Ph.D. thesis, Faculty of Engineering, Science and Mathemat-
influence of ballast degradation on its shear behaviour.” Int. J. Rail ics, Univ. of Southampton.
Transp. 6 (3): 1–18. https://doi.org/10.1080/23248378.2017.1411212. Leshchinsky, B., and H. Ling. 2013. “Effects of geocell confinement on
Drucker, D. C., and W. Prager. 1952. “Soil mechanics and plastic analysis strength and deformation behavior of gravel.” J. Geotech. Geoenviron.
or limit design.” Q. Appl. Math. 10 (2): 157–165. https://doi.org/10 Eng. 139 (2): 340–352. https://doi.org/10.1061/(ASCE)GT.1943-5606
.1090/qam/48291. .0000757.
Esveld, C. 2001. Modern railway track. Zaltbommel, Netherlands: Li, D., J. P. Hyslip, T. R. Sussmann, and S. M. Chrismer. 2015. Railway
MRT-Production. geotechnics. Boca Raton, FL: CRC.
Ferreira, F. B., and B. Indraratna. 2017. “Deformation and degradation
Downloaded from ascelibrary.org by "Indian Institute of Technology, Patna" on 04/01/21. Copyright ASCE. For personal use only; all rights reserved.

Li, D., and E. Selig. 1998a. “Method for railroad track foundation design. I:
response of railway ballast under impact loading: Effect of artificial in- Development.” J. Geotech. Geoenviron. Eng. 124 (4): 316–322. https://
clusions.” In Proc., 1st Int. Conf. on Rail Transportation, Reston, VA: doi.org/10.1061/(ASCE)1090-0241(1998)124:4(316).
ASCE. Li, D., and E. Selig. 1998b. “Method for railroad track foundation design.
Indraratna, B., J. Lackenby, and D. Christie. 2005. “Effect of confining II: Applications.” J. Geotech. Geoenviron. Eng. 124 (4): 323–329.
pressure on the degradation of ballast under cyclic loading.” Géotech- https://doi.org/10.1061/(ASCE)1090-0241(1998)124:4(323).
nique 55 (4): 325–328. https://doi.org/10.1680/geot.2005.55.4.325. Loy, H. 2008. “Under sleeper pads: Improving track quality while reducing
Indraratna, B., S. K. Navaratnarajah, S. Nimbalkar, and C. Rujikiatkamjorn. operational costs.” Eur. Railway Rev. 4: 46–51.
2014a. “Use of shock mats for enhanced stability of railroad track foun- Loy, H. 2009. “Under sleeper pads in turnouts.” Railway Tech. Rev. 2:
dation.” Aust. Geomech. J. 49 (4): 101–111. 35–38.
Indraratna, B., N. Ngo, and C. Rujikiatkamjorn. 2013. “Deformation of Marschnig, S., and P. Veit. 2011. “Making a case for under-sleeper pads.”
coal fouled ballast stabilized with geogrid under cyclic load.” J. Geo- Int. Railway J. 51 (1): 27–29.
tech. Geoenviron. Eng. 139 (8): 1275–1289. https://doi.org/10.1061
Martins, J., A. G. Correia, L. F. Ramos, J. Marcelino, L. Caldeira, and
/(ASCE)GT.1943-5606.0000864.
J. Delgado. 2009. “Measurement of vibrations induced by high-speed
Indraratna, B., N. T. Ngo, and C. Rujikiatkamjorn. 2011a. “Behavior
trains.” In Proc., Bearing Capacity of Roads, Railways and Airfields.
of geogrid-reinforced ballast under various levels of fouling.” Geotext.
8th Int. Conf. (BCR2A’09), 1311–1319. Boca Raton, FL: CRC Press.
Geomembr. 29 (3): 313–322. https://doi.org/10.1016/j.geotexmem
Navaratnarajah, S. K., and B. Indraratna. 2017. “Use of rubber mats to
.2011.01.015.
improve the deformation and degradation behavior of rail ballast under
Indraratna, B., and S. Nimbalkar. 2013. “Stress-strain degradation re-
cyclic loading.” J. Geotech. Geoenviron. Eng. 143 (6): 04017015.
sponse of railway ballast stabilized with geosynthetics.” J. Geotech.
https://doi.org/10.1061/(ASCE)GT.1943-5606.0001669.
Geoenviron. Eng. 139 (5): 684–700. https://doi.org/10.1061/(ASCE)
Ngo, N. T. 2012. “Performance of geogrids stabilised fouled ballast in rail
GT.1943-5606.0000758.
tracks.” Ph.D. thesis, Faculty of Engineering and Information Sciences,
Indraratna, B., S. Nimbalkar, and T. Neville. 2014b. “Performance assess-
Univ. of Wollongong.
ment of reinforced ballasted rail track.” Proc. Inst. Civ. Eng. Ground
Improv. 167 (1): 24–34. https://doi.org/10.1680/grim.13.00018. Ngo, N. T., B. Indraratna, and C. Rujikiatkamjorn. 2017a. “Simulation
Indraratna, B., S. Nimbalkar, and C. Rujikiatkamjorn. 2012. “Performance ballasted track behavior: Numerical treatment and field application.”
evaluation of shock mats and synthetic grids in the improvement of rail Int. J. Geomech. 17 (6): 04016130. https://doi.org/10.1061/(ASCE)GM
ballast.” In Proc., 2nd Int. Conf. on Transportation Geotechnics .1943-5622.0000831.
(ICTG), edited by S. Miura, T. Ishikawa, N. Yoshida, Y. Hisari, and Ngo, N. T., B. Indraratna, and C. Rujikiatkamjorn. 2017b. “Stabilization
N. Abe, 47–62. London: CRC Press. of track substructure with geo-inclusions: Experimental evidence and
Indraratna, B., W. Salim, and C. Rujikiatkamjorn. 2011b. Advanced DEM simulation.” Int. J. Rail Transp. 5 (2): 63–86. https://doi.org/10
rail geotechnology: Ballasted track. Rotterdam, Netherlands: CRC .1080/23248378.2017.1279085.
Press/Balkema. Nimbalkar, S., B. Indraratna, S. Dash, and D. Christie. 2012. “Improved
Indraratna, B., Q. Sun, N. T. Ngo, and C. Rujikiatkamjorn. 2017. “Current performance of railway ballast under impact loads using shock mats.”
research into ballasted rail tracks: Model tests and their practical impli- J. Geotech. Geoenviron. Eng. 138 (3): 281–294. https://doi.org/10.1061
cations.” Aust. J. Struct. Eng. 18 (3): 204–220. https://doi.org/10.1080 /(ASCE)GT.1943-5606.0000598.
/13287982.2017.1359398. Oh, J. 2013. “Parametric study on geogrid-reinforced track substructure.”
Ito, M., and K. Nagai. 2008. “Degradation issues of polymer materials used Int. J. Railway 6 (2): 59–63. https://doi.org/10.7782/IJR.2013.6.2.059.
in railway field.” Polym. Degrad. Stab. 93 (10): 1723–1735. https://doi Sol-Sanchez, M., F. Moreno-Navarro, and M. C. Rubio-Gámez. 2014.
.org/10.1016/j.polymdegradstab.2008.07.011. “Viability of using end-of-life tire pads as under sleeper pads in
Kaewunruen, S. 2007. “Experimental and numerical studies for evaluating railway.” Constr. Build. Mater. 64: 150–156. https://doi.org/10.1016/j
dynamic behaviour of prestressed concrete sleepers subject to severe .conbuildmat.2014.04.013.
impact loading.” Ph.D. thesis, Faculty of Engineering and Information Sol-Sanchez, M., F. Moreno-Navarro, and M. C. Rubio-Gámez. 2015.
Sciences, Univ. of Wollongong. “The use of elastic elements in railway tracks: A state of the art
Kerr, A. D. 2003. Fundamentals of railway track engineering. Omaha, NE: review.” Constr. Build. Mater. 75: 293–305. https://doi.org/10.1016/j
Simmons-Boardman. .conbuildmat.2014.11.027.
Lackenby, J., B. Indraratna, G. McDowell, and D. Christie. 2007. “Effect of Sun, Q. 2015. “An elastoplastic strain-based approach for analysing the
confining pressure on ballast degradation and deformation under cyclic behaviour of ballasted rail track.” Ph.D. thesis, Faculty of Engineering
triaxial loading.” Géotechnique 57 (6): 527–536. https://doi.org/10 and Information Sciences, Univ. of Wollongong.
.1680/geot.2007.57.6.527. Witt, S. 2008. The influence of under sleeper pads on railway track dynam-
Lakuši, S., M. Ahac, and I. Haladin. 2010. “Experimental investigation of ics. Linköping, Sweden: Institute of Technology, Linköping Univ.
railway track with under sleeper pad.” In Proc., 10th Slovenian Road Zhai, W. M., K. Y. Wang, and J. H. Lin. 2004. “Modelling and experiment
and Transportation Congress, 386–393. Portoroz, Slovenia: Slovenian of railway ballast vibrations.” J. Sound Vibr. 270 (4–5): 673–683.
congress on roads and transportation. https://doi.org/10.1016/S0022-460X(03)00186-X.

© ASCE 04018068-16 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(9): 04018068

You might also like