Vous êtes sur la page 1sur 11

Efficiency of Open and Infill Trenches in Mitigating

Ground-Borne Vibrations
Tulika Bose 1; Deepankar Choudhury, M.ASCE 2; Julian Sprengel 3; and Martin Ziegler 4

Abstract: In the present-day context, man-made sources of ground-borne vibration are rising at a very rapid rate due to increasing construction
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

work, blasting activities, and rapidly expanding rail and road traffic systems. As a consequence, amplified levels of ground-borne vibration
occur, causing annoyance to residents living in nearby areas, posing a threat to the stability of old structures, and interfering with instrumen-
tation works in industries. This paper discusses an investigation into the use of trenches as a means of mitigating ground vibration caused by
propagation of surface (Rayleigh) waves. Two- and three-dimensional (2D and 3D) finite-element models were developed using PLAXIS for
identifying key factors affecting the vibration isolation efficiency of open and infill trenches. Parametric studies were carried out, and the results
were analyzed to arrive at optimum values of geometrical and material properties of trenches. Numerical analysis showed that, for open
trenches, normalized depth is the decisive factor and width is of importance in trenches that are very shallow. For infill trenches, it was observed
that low-density materials perform exceedingly well as infill materials but their performance is highly sensitive to the relative shear-wave
velocity between the infill material and the in situ soil. Finally, an in-depth analysis was carried out to investigate the performance of
polyurethane foam trenches in mitigating vibrations caused by harmonic loads. The analysis was extended to study the effectiveness of
these geofoam barriers in damping out the vibrations generated by a moving train. In this case, barrier efficiency was shown to increase
with increasing train speed. The key findings suggest that trenches are a simple and effective solution for reducing ground-borne vibrations.
DOI: 10.1061/(ASCE)GT.1943-5606.0001915. © 2018 American Society of Civil Engineers.
Author keywords: Geofoam trenches; Infill trenches; Moving loads; Numerical finite-element model; Open trenches; Vibration isolation;
Wave barriers.

Introduction In addition, these waves attenuate with distance in a rather slow


manner when compared with body waves, which predominate near
Ground-borne vibrations generated from machines, construction the source of vibration. Hence, vibration-induced damages and
activities, and transportation, are often of intense levels, posing distress to structures on the surface are extremely high because of
a great challenge for engineers to build structures in such areas that the Rayleigh waves (Choudhury et al. 2014).
are serviceable to residents. The problem has become acute with the Vibration isolation using trenches (open or infill) as wave barriers
advent of high-speed railway networks that are expanding at a very may be a quick, simple, and cost-competitive way to deal with this
rapid rate across the globe. As train speeds and axle loads keep on problem. Trenches function as wave barriers by curbing the motion
increasing, vibration levels are amplified to a great extent. Energy of the traveling wave, leading to energy degeneration. An open
from the surface sources of vibration mainly propagates in the form trench acts like a finite discontinuity in the ground surface across
of Rayleigh waves, which are confined to a narrow zone near the which no energy is transmitted. For an infill trench, there is a differ-
surface of the elastic half-space (Choudhury and Katdare 2013). ence in material impedance at the junction of the in situ soil and the
trench. This causes energy redistribution across the trench in the
1
Ph.D. Research Scholar, Dept. of Civil Engineering, Technical Univ. of form of reflected and transmitted waves. Trenches are used as wave
Denmark, 2800 Kgs. Lyngby, Denmark; formerly, PG Student, Dept. of barriers in two scenarios: (1) active or near-field isolation and (2) pas-
Civil Engineering, Indian Institute of Technology Bombay, Mumbai sive or far-field isolation. In the former, they are built enclosing the
400076, India. Email: bosetulika11@gmail.com source of vibration (e.g., vibrating machines); in the latter, they are
2
Institute Chair Professor, Dept. of Civil Engineering, Indian Institute of built near the objects to be shielded (e.g., buildings to be protected
Technology Bombay, Powai, Mumbai 400076, India; Adjunct Professor,
Academy of Scientific and Innovative Research, CSIR Campus, Chennai,
from vibrations of nearby rail and road traffic).
Tamil Nadu 600113, India (corresponding author). ORCID: https://orcid In vibration isolation studies involving wave barriers, numerical
.org/0000-0002-2331-7049. Email: dc@civil.iitb.ac.in methods of analysis have found popularity. Theoretical solutions are
3
Ph.D. Research Scholar, Dept. of Geotechnical Engineering, RWTH limited, involving simplified geometries, and full-scale testing meth-
Aachen Univ., Aachen 52074, Germany. Email: sprengel@geotechnik ods are often expensive. The earliest experimental work was carried
.rwth-aachen.de out by Barkan (1962), Neumeuer (1963), and McNeil et al. (1965).
4
Professor and Director, Geotechnical Engineering and Institute of Not all of these attempts proved to be successful, but their findings
Foundation Engineering, Soil Mechanics, Rock Mechanics and Waterways gave insight into the mechanism of screening by wave barriers.
Construction, RWTH Aachen Univ., Aachen 52074, Germany. Email:
Woods (1968) performed a series of field experiments for both near-
ziegler@geotechnik.rwth-aachen.de
Note. This manuscript was submitted on July 28, 2017; approved on
and far-field isolation using open trenches. Minimum trench depths
February 7, 2018; published online on May 30, 2018. Discussion period of 0.6LR (LR = Rayleigh wavelength) for active isolation and
open until October 30, 2018; separate discussions must be submitted for 1.33LR for passive isolation were suggested, considering 75%
individual papers. This paper is part of the Journal of Geotechnical screening efficiency. Haupt (1981) carried out a number of scaled
and Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. model tests using both solid concrete barriers and lightweight bore

© ASCE 04018048-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


holes and open trenches. This study showed that barrier efficiency is domain. Most studies have focused on open trenches or selected
a function of parameters in terms of wavelength normalized dimen- infill materials like concrete or bentonite. In addition, loading
sions. Aboudi (1973), Fuyuki and Matsumoto (1980), and May and has been considered to be mostly harmonic and the trench geom-
Bolt (1982) carried out numerical studies using FEM or finite etry to be a rectangular single wall. Generalized performance of
difference method (FDM). Later, boundary element method (BEM) materials based on their characteristic properties, such as density
found popularity owing to its simplicity and was widely used in or stiffness, is missing in the literature. There is also a need for
wave propagation problems (e.g., Emad and Manolis 1985; Beskos determining the sensitivity of trench efficiency to changes in geo-
et al. 1986; Leung et al. 1987). metrical parameters. In this study, an attempt was made to fill in the
An extensive parametric study was carried out by Ahmad and gaps in the literature by studying the performance of trenches
Al-Hussaini (1991) on the use of open and infill trenches as wave methodically, beginning with the simplest case of open trenches.
barriers. The screening efficiency of open trenches was found to be Next, parametric studies were carried out for a wide range of ma-
dependent mainly on depth; that of infill trenches was reported to be terials in an infill trench for a given soil domain, and the relative
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

a function of both depth and width. Al-Hussaini and Ahmad (1991) effectiveness of these materials was determined. The efficiency of a
studied the horizontal screening efficiency of barriers and reported geo-foam barrier system was subsequently investigated in depth for
that trenches were more effective in damping vertical vibrations than both a harmonic load and a moving train. The focus of this study
horizontal vibrations. Ahmad et al. (1996) used 3D BEM to study was the performance of trenches given various materials, geomet-
active isolation of machines using open trenches; Al-Hussaini and rical parameters, system configurations, and loading applications.
Ahmad (1996) used infill trenches for the same purpose. The results Two- and three-dimensional numerical finite-element models
were compared with experiments and good agreement was found. were developed using PLAXIS, validated using results from the
Yang and Hung (1997) developed a finite-element model with in- literature, and then used to investigate open and infill trenches.
finite elements to investigate the efficiency of open and infill Parametric variations in material and geometrical properties of infill
trenches for vibrations due to passing trains. It was reported that trenches were analyzed, and the results, in terms of efficiency, were
trenches were less effective in screening low-frequency vibration. compared. Optimum barrier dimensions were also investigated. In
Hung et al. (2004) carried out similar studies and observed that this study, the soil was considered to be elastic, homogenous,
trenches were more effective in screening waves caused by a train and isotropic. The loading considered was initially periodic and
moving at supercritical speed compared with subcritical speed. harmonic and later modified to simulate a moving train.
Ju and Lin (2004) reported similar results.
Adam and Estorff (2005) employed coupled BEM-FEM in a time
domain to study the effectiveness of trenches in reducing building Vibration Isolation Efficiency of Open Trenches
vibrations. They found that 80% of the forces in the building com-
ponent could be reduced by a well-designed barrier. Andersen and
Nielsen (2005) applied the same coupled approach and reported that Numerical Model
trenches were better at mitigating vertical vibrations than horizontal A numerical model was developed to understand the behavior and
vibrations. Wang et al. (2006, 2009) numerically investigated the efficiency of open trenches, as wave barriers, in mitigating ground-
efficiency of expanded polystyrene (EPS) barriers in protecting borne vibrations generated by a harmonic load vibrating in the ver-
buried structures under blast load. Leonardi and Buonsanti tical direction. The 2D axisymmetric model consisted of 15 noded
(2014) studied the efficiency of concrete and compacted soil barriers triangular elements. Average element size was fixed, based on the
for reducing train-induced vibrations. Esmaeili et al. (2014) and recommendations of Kuhlemeyer and Lysmer (1973), to be one-
Zakeri et al. (2014) investigated V-shaped and step-shaped trenches, eighth to one-tenth of the wavelength. In order to account for
respectively. Their findings revealed that trenches with such modi- the semi-infinite extent of the soil, viscous boundary conditions
fied geometries were more effective than conventional rectangular were assigned along the model edges to avoid undue wave reflec-
trenches. tions. Standard fixities were applied, wherein the vertical sides
Full-scale experimental studies were conducted by Massarsch were restrained horizontally (ux ¼ 0) and the bottom was fully re-
(1991) on the efficiency of gas cushion screen systems. These sys- strained (ux ¼ uy ¼ 0). A linear elastic soil model was chosen be-
tems were found to be comparable to open trenches. Baker (1994) cause wave propagation in soil involving trenches usually generates
carried out field tests on stiffer and softer barriers made of concrete small strains, meaning that material nonlinearities arising from
and bentonite, respectively. Davies (1994) carried out 20-g centri- small variations in stress over a cycle are not very influential.
fuge tests to study the screening effectiveness of EPS barriers on
buried objects. The tests indicated that low acoustic materials could
reduce the magnitude of ground shock loading on buried structures.
Zeng et al. (2001) performed tests on rubber-modified asphalt and
found that, owing to a high damping ratio, the modified asphalt
could be used effectively beneath high-speed railway tracks as a
foundation material for vibration attenuation. Itoh et al. (2005) con-
ducted centrifuge tests and suggested using a combination of crumb
rubber–modified asphalt at the vibration source and an EPS barrier
along the transmission path. Murillo et al. (2009) performed 50-g
centrifuge tests on EPS barriers and reported incremental efficiency
to be a function of barrier depth. Alzawi and EI-Naggar (2011)
carried out full-scale field tests to study the effectiveness of geofoam
barriers. Their findings revealed a significant increase in perfor-
mance for normalized barrier depth greater than 0.6.
Research to date has lacked a systematic procedure for selecting
Fig. 1. Schematic of vibration isolation system using an open trench.
the best infill material to be used in trenches for a given soil

© ASCE 04018048-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Typical 2D numerical model developed for an open trench in PLAXIS.

Considering this, soil behavior at small strain levels were assumed trench, particle displacements were fairly insignificant and the influ-
to be linearly elastic without significant loss of accuracy. (Yang and ence of the trench almost disappeared. To enumerate Ar, then, x ¼
Hung 1997; Andersen and Nielsen 2005; Alzawi and EI-Naggar 10LR was used in this study, as in the studies carried out by Ahmad
2011). In wave propagation problems involving barriers, in order and Al-Hussaini (1991) and Yang and Hung (1997).
to prevent the results from depending on load frequency, trench
geometrical parameters are usually normalized with those of the
Rayleigh wavelength, LR (Ahmad and Al-Hussaini 1991). Fig. 1 is Parametric Study
a schematic of the developed numerical model; Fig. 2 shows the The three variables that define an open trench are depth, d, width,
model’s meshing details. w, and screening distance l, which can be optimized to achieve
maximum screening efficiency. In this study, the trench was placed
Model Validation at different locations. Different combinations of width and depth
The results of any vibration isolation scheme are typically ex- were chosen against which system efficiency could be evaluated.
pressed as an amplitude reduction ratio, ARR (Woods 1968), which The input parameters for the soil domain were the same as
is given as those in Yang and Hung (1997). The relevant properties were
density, ρ ¼ 1,800 kg=m3 ; shear-wave velocity, V S ¼ 101 m=s;
AI
ARR ¼ ð1Þ Rayleigh wave velocity, V R ¼ 93 m=s; LR ¼ 3 m; Poisson’s ratio,
AO ν ¼ 0.25; and damping coefficient, ξ ¼ 5%. The source of vibration
was taken to be a periodic harmonic load of magnitude 1 kN vibrat-
where AI = displacement or velocity amplitude after trench instal-
ing vertically at a frequency of 31 Hz. For practical purposes, the
lation; and AO = displacement or velocity amplitude before trench
footing carrying the vibrating load was not included in the numerical
installation.
model because it did not alter or affect the results of the study
The values of ARR vary at different locations beyond the trench.
(Kattis et al. 1999).
To have an idea of the overall performance of the barrier, an average
Fig. 4 shows the results of the parametric study, plotted in terms
was computed by integrating ARR over the barrier influence zone,
of variation in average amplitude reduction ratio/system efficiency
x, represented by the average amplitude reduction ratio (Ar)
with changes in the trench’s normalized geometrical parameters.
Z The barrier was placed at two screening distances (L ¼ 3 and
1
Ar ¼ ðARRÞdx ð2Þ L ¼ 5) and analyzed for a wide range of depth, D, and width
x
W, values. From the figure, it can be observed that open trenches
From this, overall system efficiency or effectiveness (Ef) could have excellent vibration isolation capacity. In the range considered
be evaluated as for the parametric study here, the minimum efficiency of the system
was as high as 55% whereas the maximum was more than 80%. It
Ef ¼ ð1 − ArÞ × 100 ð3Þ became evident that normalized depth was the key parameter con-
trolling system effectiveness, which was maximized with increas-
The numerical model was first validated, as mentioned previ- ing normalized depth, D. This was true for all trench locations and
ously. For that purpose, an open trench of depth d ¼ 1.0LR , width widths. For an efficiency Ef > 60%, D had to be greater than 0.8.
w ¼ 0.1LR , and screening distance l ¼ 5LR was considered. In addition, the performance of the trench was observed not to be
Fig. 3(a) plots the variation in ARR with normalized distance be- very sensitive to the barrier location, L. The same was true for the
yond the source of vibration. Good agreement was found between trench width, W, with the exception of very shallow trenches. For
the simulated results and those reported in the literature. these cases, the response of the system improved with increasing
To compute system efficiency as a whole, average amplitude re- width, mainly because open trenches represent a discontinuity in
duction was calculated over a zone of influence, x, beyond the the ground profile across which no part of the wave energy is al-
barrier. For this purpose, the area over which the trench exerted lowed to pass, meaning that wave reflection plays the major role.
its influence was determined by plotting the normalized soil particle Hence, for a sufficiently deep barrier that obstructs Rayleigh waves,
displacement beyond the barrier, as shown in Fig. 3(b). It is evident creation of a finite discontinuity in the ground surface is sufficient.
from this plot that, after a distance of roughly 10LR beyond the open However, for a very shallow trench, as in this case, not all of the

© ASCE 04018048-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


2.5 1.2
Present study Open Trench
Ahmad and Al- 1.0 No Trench
2.0 Hussaini 1991
Beskos et al. 1986
0.8
1.5
0.6
1.0
0.4

0.5
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

0.2

0.0 0.0
0 2 4 6 8 10 12 0 5 10 15
(a) (b)

Fig. 3. Analysis of an open trench: (a) 2D finite-element model verification, W ¼ 0.1, D ¼ 1, L ¼ 5; and (b) normalized vertical displacement
amplitude of the ground surface.

0.5 50
L=3,W=0.17
L=3,W=0.33
0.4 60
L=3,W=0.50
L=5,W=0.17
0.3 L=5,W=0.33 70
L=5,W=0.50

0.2 80

0.1 90

0 100
0.5 1.0 1.5

Fig. 4. Variation in the average amplitude reduction ratio with changes


in the normalized depth of an open trench.

Fig. 5. Variation in the average amplitude reduction ratio with changes


wave energy is blocked by the barrier, meaning that width has an in the shear-wave velocity ratio of an infill trench and in situ soil.
important role to play. This observation is consistent with the find-
ings reported by Ahmad and Al-Hussaini (1991).

considered for the soil domain were ρsoil ¼ 1,850 kg=m3 , V Ssoil ¼
Vibration Isolation Efficiency of Infill Trenches 225 m=s, ν soil ¼ 0.4, and ξ soil ¼ 5%. The dynamic load was
simulated to be periodic and harmonic, vibrating vertically at a fre-
Two-Dimensional Parametric Study quency of 45 Hz. The barrier was placed at a fixed distance of 2.5 m
from the load and was of a constant depth of 3 m (D ¼ 0.65) and a
Open trenches, though an excellent method of mitigating ground-
width of 0.25 m. The ratio of infill material density to soil density,
borne vibrations, find their use in limited cases owing to stability
ρfill =ρsoil , was varied from 0.02 to 4.20. The shear-wave velocity
issues. Hence, infill trenches become a popular choice when the
ratio of the infill material to that of the soil, V Sfill =V Ssoil , was
wavelength exceeds a depth beyond which open vertical cuts find
changed from 0.25 to 6.0. The damping properties of the infill were
difficulty in construction and stability. For an open trench, wave
kept in the 5–15% range.
reflection plays the major role, whereas, for an infill trench, the
combination of energy in the reflected and transmitted waves
governs efficiency. Results and Discussion
In this study, infill materials having both lower and higher den-
sities (ρfill ) compared with the soil domain were chosen. For each Fig. 5 plots the variation in Ar with changes in ratio of infill shear-
material density, the shear-wave velocity (V Sfill ) was gradually in- wave velocity to that of the soil for various material density ratios. It
creased from low to high. A parametric study was carried out with is observed that the functioning of the trench was largely dependent
these widespread spectra of infill materials to assess their relative on the contrast between the properties of the infill and those of the in
efficiencies. Also, for a few chosen densities the damping charac- situ soil. Infill density and shear-wave velocity had a great impact on
teristics of the materials were varied. The relevant parameters the results, which were different for low- and high-density materials.

© ASCE 04018048-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


First, the impact of geofoam trench depth, D, and screening
distance, L, were analyzed while keeping trench width, w,
constant at 0.25 m. Following that, w was varied for chosen depths,
keeping the trench fixed at two locations simulating near-field
and far-field isolation. Finally, the influence of cross-sectional area,
A, and ratio, d=w, was investigated. Soil and loading parameters
were the same as those for the infill trenches; the properties of
the PU foam, taken from Alzawi and EI-Naggar (2011), were
V S ¼ 330 m=s, ρ ¼ 61 kg=m3 .

Results and Discussion


Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

Influence of Barrier Depth and Location


The trench was placed at different locations, L, at which the nor-
malized depth, D, was changed from 0.3 to 1.5; the width was kept
Fig. 6. Variation in the average amplitude reduction ratio with changes
constant. Fig. 7(a) shows the combined influence of D and L on the
in the damping characteristics of an infill trench.
average amplitude reduction ratio, Ar. First, it was observed that the
PU foam trenches had very good vibration isolation capacity; Ar
showed a major dependency on both screening distance and normal-
In general, low-density materials (ρfill =ρsoil < 0.15) performed ized depth. Second, it was observed that on changing the normalized
well as wave barriers compared with high-density materials. In fact, distance, L, from 0.4 to 1.6, Ar changed in a complex manner de-
their performance was comparable to that of open trench. However, pending on depth. For L > 1.8, effectiveness was mostly governed
their response depended on the relative shear-wave velocity of the by normalized depth and was almost independent of the barrier’s
infill material and the in situ soil, V Sfill =V Ssoil . The system per- distance from the source of vibration. Thus, it might be said that
formed efficiently, displaying lower Ar values, when the shear- increasing D generally results in decreasing Ar, but in a complicated
wave velocity of the fill material was lower than that of the soil way that depends on barrier position. For far-field isolation, an in-
because, in this case, the low-density materials had sufficient en- crease in D was generally accompanied by a boost in system effi-
ergy dissipation capacity. With increasing infill shear-wave veloc- ciency; for near-field isolation, the same was not always true. Third,
ity, either Ar increased or efficiency decreased. An upper limiting increasing the barrier’s depth, D, beyond 1.1 or 1.2 did not have any
value could be identified as 1; as V Sfill =V Ssoil approached 1.0, Ar significant impact on Ar. The optimum barrier depth could therefore
increased sharply. be considered approximately 1.2 for all practical purposes.
On the other hand, dense infill materials (ρfill =ρsoil > 1.0) also
Influence of Barrier Width
functioned effectively in the trench, exhibiting lower Ar at very high
Geofoam trenches were placed at two locations—in the first case
shear-wave velocities compared with the in situ soil. The lower limit
simulating relatively near-field isolation, with L ¼ 0.4, and in the
in this case was 2.5. For V Sfill =V Ssoil > 2.5, Ar was generally lower
second case simulating far-field isolation, with L ¼ 1.5. In both
or efficiency was higher, indicating that high-density materials
cases, simulations were performed for a few chosen depths by vary-
having sufficient stiffness could resist the incoming wave. Materials
ing the width of the trench and keeping the barrier location fixed.
having a density of 0.15 < ρfill =ρsoil < 1 performed well when the
Figs. 7(b and c) show the influence of width, w, on average ampli-
shear-wave velocity was 1.0 > V Sfill =V Ssoil > 2.5).
tude reduction ratio. It can be clearly seen that, unlike in the open
Fig. 6 shows the variation in Ar with changing infill material
trench, where width did not play a significant role in system effi-
damping keeping all other parameters unchanged. It is observed
ciency, in the geofoam trenches width had a substantial impact
that Ar was not very sensitive to changes in the damping character-
on performance. In an infill trench, wave reflection, absorption,
istics of the infill materials. When the damping coefficient of the
and transmission all have a role to play and hence overall system
infill increased from 5 to 15%, no significant changes were detected
stiffness is important, in which width is an integral part. In this study,
in Ar.
with an increase in width from 0.15 to 0.35, Ar decreased by
approximately 40% for almost all depths and for both isolation
cases. Fig. 7(b) shows that, for near-field isolation increasing depth
Vibration Isolation Efficiency of Geofoam Trenches did not have the same impact on system efficiency that it did for
far-field isolation [Fig. 7(c)].
The results described in the previous section indicate that low-
density materials having lower shear-wave velocity relative to the Influence of Barrier Cross-Sectional Area and Slenderness
surrounding soil are ideal for use as trench infill. Additional analyses Ratio
were carried out on a low-density geofoam infill material. The type For the barrier at L ¼ 1.5, the influence of cross-sectional area, A,
of geofoam chosen was polyurethane, a leading member of the on performance was investigated along with the optimum d=w ratio
wide ranging and diverse family of polymers. Polyurethane is manu- for each cross-sectional area. The size of the cross-sectional area
factured in both solid and cellular forms and can be rigid as well as was increased gradually from 0.5 to 2.5 m2 . For each area, the slen-
flexible. derness of the trench (d=w) was progressively varied from 0.5 to
approximately 6.0. Fig. 7(d) shows the combined influence of
cross-sectional area and slenderness ratio on the functioning of
Two-Dimensional Parametric study
the trenches. For smaller cross-sectional areas (A < 1.0 m2 ), it
A stepwise sensitivity analysis was carried out to investigate how was reasonable to construct a deep trench (d=w ¼ 4.0–5.0 m)
the geometrical parameters of polyurethane foam trench (PU foam) rather than a shallow one for greater efficiency. However, for larger
influence the efficiency. cross-sectional areas (A > 1.0 m2 ), it was sufficient to construct a

© ASCE 04018048-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


1.4 0.7
D=0.3 D= 0.60
D=0.5 D= 0.80
1.2 D=0.8 0.6
D= 1.00
D=1.0
D=1.2
1.0 D=1.5 0.5

0.8 0.4

0.6 0.3

0.4 0.2
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

0.5 1.0 1.5 2.0 2.5 0.1 0.2 0.3 0.4 0.5
(a) (b)

0.9 1.2
D= 0.60 A=0.5
0.8 D= 0.80 A=0.8
1.0 A=1.2
D= 1.00
A=1.5
0.7 A=2.0
0.8 A=2.5
0.6
0.6
0.5

0.4
0.4

0.3 0.2
0.1 0.2 0.3 0.4 0.5 0 1 2 3 4 5 6
(c) (d)

Fig. 7. Variation in the average amplitude reduction ratio with changes in various geometrical parameters of a geofoam trench: (a) normalized
screening distance; (b and c) width; and (d) slenderness ratio.

shallow trench (d=w ¼ 1.5–2.0 m). Extra cost incurred in creating present study. The boundary conditions were to (1) completely re-
deeper trenches does not bring about greater benefits. In fact, with strain the bottom edge and (2) prevent the vertical model boundaries
an increasing cross-sectional area, the optimum d=w ratio hovered from moving in the direction of their normal. Element size was kept
near 1.5. Again, for low slenderness ratio values, d=w < 2, an in- roughly less than one-eighth of the smallest Rayleigh wavelength
creasing cross-sectional area had a very positive impact on system (Kuhlemeyer and Lysmer 1973). In addition, local refinement of
efficiency. For d=w ¼ 1, Ar decreased by approximately 55% when the mesh was done near the critical areas of interest such as the load-
A increased from 0.5 to 2.5 m2 . For d=w > 2, increasing area did ing zone and the barrier location and in general on the ground sur-
not have much of an impact on performance except for very small face to ensure accuracy of the results. Fig. 8(a) shows the discretized
cross sections (A < 0.80 m2 ). 3D model developed for this problem. A linear elastic model was
adopted for all materials considering a small strain behavior.
The numerical model was first validated with results from the
Three-Dimensional Finite-Element Model and literature. For this purpose, the field data recorded by Alzawi and
Analysis EI-Naggar (2011) were taken for comparison. Fig. 8(b) shows good
agreement between the results obtained in this study and those ob-
Validation of Present Model served by Alzawi and EI-Naggar (2011). The differences in some
cases could be due to variability and anisotropy in soil properties in
After an extensive 2D parametric study of the performance of PU localized areas in the field.
foam trenches as wave barriers, a 3D analysis of their responses
with other configurations and loading conditions was carried
out. A 3D finite-element model (100 × 50 × 20 m) was developed Results and Discussion
in the PLAXIS 3D dynamic module using 10 noded tetrahedral
elements. The model dimensions were chosen to avoid any boun- Influence of Barrier System Type
dary effects (Kumar et al. 2017; Kumar and Choudhury 2018). The 3D model was next employed to study the response of other
Viscous boundaries were applied along the edges to account for geofoam barrier configurations. Usually, the most common profile
the semi-infinite extent of the soil and to prevent undue wave adopted for wave barriers is a straight, rectangular vertical cut into
reflection along the boundaries (Kumar et al. 2015, 2016). Studies the ground. In this section, calculations are performed with another
have shown that the wave relaxation coefficients related to absorb- simple barrier configuration: two continuous foam walls kept at
ent boundaries, C1 ¼ 1.0 and C2 ¼ 0.25, result in reasonably good the spacing, s, shown in Fig. 9(a). Simulations were carried out
absorption of waves at the edges (Wang et al. 2009; Brinkgreve and to determine the influence of s between the two walls on system
Vermeer 1998). Accordingly, these coefficients were adopted in the performance. Analyses were performed by varying the normalized

© ASCE 04018048-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


(a) (b)
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. 3D analysis of geofoam trenches: (a) typical model developed in PLAXIS; and (b) validation of the numerical model, l ¼ 2.5 m, f ¼ 50 Hz.

(a) (b)

Fig. 9. Analysis of a double-walled continuous rectangular geofoam trench system: (a) schematic; and (b) variation in the average amplitude
reduction ratio with changes in the normalized spacing between walls.

spacing, S ¼ s=LR , from 0.2 to 1.0 for a 30–60-Hz frequency range. Fig. 10 shows the developed numerical model, which had
The PU foam barriers were of normalized depth, D ¼ 0.75 m, and dimensions of 200 × 100 × 20 m—large enough to prevent wave
width, w ¼ 0.2 m, and were placed at location, L ¼ 0.53 m, from reflections from the boundaries. The track rested on an embank-
the source of vibration. The size, fixities, boundary conditions, and ment of width 5 m and height 0.5 m. For simplicity, the properties
meshing of the 3D model were the same as described in the preced- of the soil in the embankment and the ground remained the same
ing section. The material properties of the in situ soil and geofoam as described previously. The track consisted of a pair of steel rails
were also unchanged. The values of Ar were computed for vertical resting on concrete sleepers, both modeled using beam elements.
velocity by observing the time history of nodes on the ground The rail and sleeper cross-sectional areas were, respectively, Arail ¼
surface along a monitoring path. Fig. 9(b) shows the variation in 0.0077 m2 and Asleeper ¼ 0.05 m2 . The sleepers were laid on the
average amplitude reduction ratio as a function of barrier normalized ground at a spacing, c, of 0.6 m.
spacing for the chosen frequency range. It can be seen that this The vehicle unit chosen for this demonstration was a typical
barrier system was quite effective in damping out vertical soil vibra- German ICE3 railcar with a distance, X, between the first and last
tions. The Ar values were much lower at higher frequencies, indi-
cating better system performance. In addition, the system response
was quite sensitive to the barrier spacing, especially at low values,
S ¼ 0.2–0.4. In the frequency range chosen for this study, the
optimum spacing was approximately 0.5–0.6LR. Wider spacing
would not bring any added benefits in terms of increasing efficiency
and could even be detrimental to the performance in some cases.

Barrier Performance for a Moving Load


Simulation of Moving Load. Up to this point, a harmonic load
was considered in all analyses. However, the scenario was changed
for a load that, apart from being dynamic, shifts its position with
time, which is the case for a moving load such as a train. Because a
moving load, compared with a static load, can significantly increase
displacements in a structure, the response of the PU foam barrier
Fig. 10. 3D numerical model developed for simulating a moving load.
under moving-load vibrations was investigated.

© ASCE 04018048-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


wagon axles of 21.6 m. The length of the loading in the rail, Xo, To incorporate moving load, dynamic signals multipliers were
was Xo ¼ X þ 2 × 0.3X ≈ 34.80 m, with the additional length assigned to each of these 117 point loads. The signal for each load
accounting for shear force distribution and impact load distribution location represented how the forces varied at that particular point in
effects (Shahraki et al. 2014). To replicate a train moving on the the rail as the wheel load moved along. The multipliers were ob-
rails, point loads were applied along the length of the beam at a tained by considering the rail to be a beam resting on an elastic pin
spacing of c=2 ¼ 0.30 m. Thus, the total number of dynamic loads foundation and analyzed under a set of unit static loads at different
(N) per rail was N ¼ Xo=c=2 ¼ 117.The value assigned to each locations. Each point load was multiplied by the value of its own
point load was the vertical wheel load (P ¼ 80 kN). signal for every dynamic time step. The latter parameter accounted
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 11. Variation in soil particle velocity with changes in train speed in the absence of a trench: (a) time domain; and (b) frequency domain.

(a) (b)

(c)

Fig. 12. Comparative analysis of velocity in the frequency domain in the presence and absence of a trench: (a) 80 km=h; (b) 180 km=h; and
(c) 250 km=h.

© ASCE 04018048-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


for the time taken by the train to cross a distance of c=2. For caused by harmonic loads. A 3D analysis was also carried out for
example, a train traveling at a speed, V, of 180 km=h would dynamic loads that change position with time.
exibit a time lag between two consecutive point loads of Observations from the results of this study are as follows:
Δt ¼ c=2=V ¼ 0.3 m=50 m=s ¼ 0.006 s. Accordingly, the time • Open trenches performed exceedingly well in mitigating ground
step for this dynamic analysis was 0.006 s. The total time of analy- vibrations. The main parameter controlling efficiency was nor-
sis was based on the time taken by the last axle to cross the loading malized depth, D. For D > 0.8, system effectiveness in all cases,
zone. In this study, analyses were carried out for train speeds of irrespective of location or width, was found to be greater than
250, 180, and 80 km=h. At these chosen speeds, track responses 75%. Trench width did not play a very important role except
could be assumed to be mainly quasi-static and the dynamic effects in extremely shallow trenches. The efficiency of an open trench
to be negligible. Dynamic forces due to wheel-rail irregularity and as a wave barrier was not very sensitive to screening distances.
defects such as wheel flats were neglected. Here, the focus was on • For infill trenches, the most important parameters governing
quasi-static track response. efficiency were density and shear-wave velocity of the infill
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

The geofoam trench was chosen to be of depth 5 m and width material relative to the in situ soil; the damping characteristics
0.5 m. It was placed at roughly 10 m from the center of the track. of the infill material were not significant. It was shown that both
The material properties of the foam and the soil domain were the low- and high-density materials (in comparison with the in situ
same as before. soil) are ideal for use in infill trenches but their performance is
Results and Discussion. Fig. 11(a) shows the influence of train highly sensitive to the relative stiffness of the trench material and
speed on the velocity of soil particles on the ground surface. It is the in situ soil. For the former category (ρfill =ρsoil < 0.15), the
seen that, with increasing train speed, vibration velocity increased, upper limit is V Sfill =V Ssoil < 1.0; for the latter (ρfill =ρsoil > 1),
especially in the near-field region. This was most notable for vertical the lower limit is V Sfill =V Ssoil > 2.5.
vibrations, which were very high in the near field, although their • PU foam trenches proved to be very effective in damping out
attenuation with distance occurred at a very fast rate. At distances ground-borne vibrations. Their efficiency was shown to depend
far away from the source of vibration, horizontal and vertical veloc- on normalized depth, width, screening distance, and, d=w ratio.
ities were nearly the same for all train speeds. Fig. 11(b) shows In areas near the source of vibration (0.4 < L < 1.8), the barrier
typical results of the analysis in the frequency domain, comparing showed a greater dependency on both screening distance and
the velocity of vibration for different speeds in the absence of a depth, but in regions far away (L > 1.8) the influence of screen-
trench. It is observed that, with increasing train speed, ground vibra- ing distance was almost eliminated. The optimum barrier depth
tion frequency increased. For a train speed of 80 km=h, vibration for all purposes could be taken as 1.2. Increasing width was
frequency ranged 0–30 Hz. For speed of 180 km=h, the predominant shown to have a positive impact on the barrier in both near-field
range of vibration was 10–40 Hz, and for a speed of 250 km=h, the and far-field isolation. Regarding cross-sectional area, for
range was 20–55 Hz. Fig. 12 compares the frequency of vertical A < 1.0 m2 a deeper trench (d=w ¼ 4.0–5.0) was optimal.
vibration, in the presence and absence of the trench, for the different However, for A > 1.0 m2 the optimal d=w ratio was 1.5–2.0.
train speeds. From Fig. 12(a), it is clearly observed that, at 80 km=h, • Three-dimensional analysis revealed that double-walled contin-
vibration frequency post-trench installation was mostly arrested uous rectangular trenches performed well as wave barriers but
within 20 Hz. Frequencies of 20–30 Hz were partly damped their functioning was sensitive to normalized spacing. Optimal
by the wave barrier. The same trend is noted in Figs. 12(b and c). normalized spacing in this study was shown to be roughly
Fig. 12(b) shows that frequencies of 30–40 Hz were mostly damped 0.5–0.6 times the Rayleigh wavelength.
out. The particles vibrated primarily at 0–30 Hz, and especially at • The barriers were found to be quite effective in damping out
10–20 Hz. In the latter case [Fig. 12(c)], frequencies higher than vibrations caused by passage of a moving load. They mostly
40 Hz were completely blocked by the barrier. This shows that damped out the high-frequency or shorter-wavelength compo-
the high-frequency or shorter wavelength waveforms were blocked nents from the vibration velocities, indicating increasing system
very effectively by the barrier. For the chosen barrier depth (5 m) and efficiency with increasing train speed.
soil profile, corresponding to a frequency of 40 Hz, the normalized These observations can be generalized to arrive at the conclu-
depth, D, was approximately 1.0. For frequencies higher than 40 Hz, sion that trenches are very effective when used as wave barriers to
D was greater than 1 and the barrier effectively depleted these mitigate ground-borne vibrations.
frequency contents. Hence, more efficiency was achieved at higher
train speeds, as in this case the quasi-static track response had higher
frequency contents. Acknowledgments
The first author would like to thank the German Academic
Exchange Service (DAAD) for providing financial assistance to
Conclusions carry out a part of this study at RWTH Aachen University in the
A numerical finite-element analysis was carried out using PLAXIS DAAD-IIT Masters Students Exchange Program.
to interpret the behavior of open and infill trenches when acting as
wave barriers to reduce ground-borne vibrations. The behavior and
responses of trenches having a wide range of geometrical and Notation
material properties, different barrier types, and loading conditions
The following symbols are used in this paper:
were analyzed. This analysis was carried out in stages, on both
open and infill trenches, with a special focus on polyurethane ARR = amplitude reduction ratio;
foam trenches. The developed model was used to identify the Ar = average amplitude reduction ratio;
key factors affecting the vibration isolation capacity of trenches. A = cross-sectional area;
Two-dimensional simulations were performed in order to under- D = normalized trench depth;
stand the impact of geometrical and material properties of the d = trench depth;
trenches on their efficiency as wave barriers against vibrations E = elastic modulus;

© ASCE 04018048-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


f = frequency; Modelling—Centrifuge, 319–324. Rotterdam, Netherlands: A.A.
G = shear modulus; Balkema.
Emad, K., and G. D. Manolis. 1985. “Shallow trenches and propagation of
L = normalized distance;
surface waves.” J. Eng. Mech. 111 (2): 279–282. https://doi.org/10
LR = Rayleigh wavelength; .1061/(ASCE)0733-9399(1985)111:2(279).
l = distance; Esmaeili, M., J. Zakeri, and S. Mosayebi. 2014. “Investigating the opti-
N = number of dynamic loads; mized open V-shaped trench performance in reduction of train-induced
P = vertical wheel load; ground vibrations.” Int. J. Geomech. 14 (3): 04014004. https://doi.org
S = normalized spacing; /10.1061/(ASCE)GM.1943-5622.0000331.
Fuyuki, M., and Y. Matsumoto. 1980. “Finite difference analysis of
s, c = spacing;
Rayleigh wave scattering at a trench.” Bull. Seismol. Soc. Am. 70 (6):
V R = Rayleigh wave velocity; 2051–2069.
V S = shear-wave velocity; Haupt, W. A. 1981. “Model tests on screening of surface waves.” In Vol. 3 of
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

W = normalized trench width; Proc., 10th Int. Conf. on Soil Mechanics and Foundation Engineering,
w = trench width; 215–222. Stockholm.
X = axle distance in railcars; Hung, H. H., Y. B. Yang, and D. W. Chang. 2004. “Wave barriers for
reduction of train-induced vibrations in soils.” J. Geotech. Geoenviron.
Δt = dynamic time step;
Eng. 130 (12): 1283–1291. https://doi.org/10.1061/(ASCE)1090-0241
ν = Poisson’s ratio; (2004)130:12(1283).
ξ = damping coefficient; and Itoh, K., X. Zeng, M. Koda, O. Murata, and O. Kusakabe. 2005. “Centrifuge
ρ = density. simulation of wave propagation due to vertical vibration on shallow
foundations and vibration attenuation countermeasures.” J. Vibr. Control
11 (6): 781–800. https://doi.org/10.1177/1077546305054150.
References Ju, S. H., and H. T. Lin. 2004. “Analysis of train-induced vibrations and
vibration reduction schemes above and below critical Rayleigh speeds
Aboudi, J. 1973. “Elastic waves in half-space with thin barrier.” J. Eng. by finite element method.” Soil Dyn. Earthquake Eng. 24 (12):
Mech. 99 (1): 69–83. 993–1002. https://doi.org/10.1016/j.soildyn.2004.05.004.
Adam, M., and O. Von Estorff. 2005. “Reduction of train-induced building Kattis, S. E., D. Polyzos, and D. E. Beskos. 1999. “Vibration isolation by a
vibrations by using open and filled trenches.” Comput. Struct. 83 (1): row of piles using a 3-D frequency domain BEM.” Int. J. Numer.
11–24. https://doi.org/10.1016/j.compstruc.2004.08.010. Methods Eng. 46 (5): 713–728. https://doi.org/10.1002/(SICI)1097
Ahmad, S., and T. M. Al-Hussaini. 1991. “Simplified design for vibration -0207(19991020)46:5%3C713::AID-NME693%3E3.0.CO;2-U.
screening by open and in-filled trenches.” J. Geotech. Eng. 117 (1): Kuhlemeyer, R. L., and J. Lysmer. 1973. “Finite element method accuracy
67–88. https://doi.org/10.1061/(ASCE)0733-9410(1991)117:1(67). for wave propagation problems.” J. Soil Mech. Found. Div. 99 (5):
Ahmad, S., T. M. Al-Hussaini, and K. L. Fishman. 1996. “Investigation on 421–427.
active isolation of machine foundations by open trenches.” J. Geotech. Kumar, A., and D. Choudhury. 2018. “Development of new prediction
Eng. 122 (6): 454–461. https://doi.org/10.1061/(ASCE)0733-9410 model for capacity of combined pile-raft foundations.” Comput.
(1996)122:6(454). Geotech. 97 (May): 62–68. https://doi.org/10.1016/j.compgeo.2017
Al-Hussaini, T. M., and S. Ahmad. 1991. “Design of wave barriers for .12.008.
reduction of horizontal ground vibration.” J. Geotech. Eng. 117 (4): Kumar, A., D. Choudhury, and R. Katzenbach. 2016. “Effect of earthquake
616–636. https://doi.org/10.1061/(ASCE)0733-9410(1991)117:4(616). on combined pile-raft foundation.” Int. J. Geomech. 16 (5): 04016013.
Al-Hussaini, T. M., and S. Ahmad. 1996. “Active isolation of machine https://doi.org/10.1061/(ASCE)GM.1943-5622.0000637.
foundations by in-filled trench barriers.” J. Geotech. Eng. 122 (4): Kumar, A., D. Choudhury, J. Shukla, and D. L. Shah. 2015. “Seismic
288–294. https://doi.org/10.1061/(ASCE)0733-9410(1996)122:4(288). design of pile foundation for oil tank by using PLAXIS3D.” Disaster
Alzawi, A., and M. H. El Naggar. 2011. “Full scale experimental study on Adv. 8 (6): 33–42.
vibration scattering using open and in-filled (geofoam) wave barriers.” Kumar, A., M. Patil, and D. Choudhury. 2017. “Soil-structure interaction in
Soil Dyn. Earthquake Eng. 31 (3): 306–317. https://doi.org/10.1016/j a combined pile-raft foundation: A case study.” Proc. Inst. Civ. Eng.
.soildyn.2010.08.010. Geotech. Eng. 170 (2): 117–128. https://doi.org/10.1680/jgeen.16
Andersen, L., and S. R. K. Nielsen. 2005. “Reduction of ground vibration
.00075.
by means of barriers or soil improvement along a railway track.” Soil
Leonardi, G., and M. Buonsanti. 2014. “Reduction of train-induced
Dyn. Earthquake Eng. 25 (7–10): 701–716. https://doi.org/10.1016/j
vibrations by using barriers.” Res. J. Appl. Sci. Eng. Technol. 7 (17):
.soildyn.2005.04.007.
623–632. https://doi.org/10.19026/rjaset.7.715.
Baker, J. M. 1994. “An experimental study on vibration screening by
Leung, K. L., I. G. Vardoulakis, and D. E. Beskos. 1987. “Vibration
in-filled trench barriers.” M.Sc. thesis, State Univ. of New York.
Barkan, D. D. 1962. Dynamics of bases and foundations, Chaps. VIII and isolation of structures from surface waves in homogeneous and nonho-
IX. 374–406. New York, NY: McGraw-Hill. mogeneous soils.” In Soil-structure interaction, edited by A. S. Cakmak,
Beskos, D. E., G. Dasgupta, and I. G. Vardoulakis. 1986. “Vibration 155–169. Amsterdam, Netherlands: Elsevier.
isolation using open or filled trenches. Part I: 2-D homogeneous soil.” Massarsch, K. R. 1991. “Ground vibration isolation using gas cushions.” In
Comput. Mech. 1 (1): 43–63. https://doi.org/10.1007/BF00298637. Proc., 2nd Int. Conf. on Recent Advances in Geotechnical Earthquake
Brinkgreve, B. J., and P. A. Vermeer. 1998. PLAXIS finite element code for Engineering and Soil Dynamics, 1461–1470. Rolla, MO: Univ. of
soil and rock analysis. Delft, Netherlands: A.A. Balkema. Missouri at Rolla.
Choudhury, D., and A. D. Katdare. 2013. “New approach to determine May, T. W., and B. A. Bolt. 1982. “The effectiveness of trenches in reduc-
seismic passive resistance on retaining walls considering seismic ing seismic motion.” Earthquake Eng. Struct. Dyn. 10 (2): 195–210.
waves.” Int. J. Geomech. 13 (6): 852–860. https://doi.org/10.1061 https://doi.org/10.1002/eqe.4290100203.
/(ASCE)GM.1943-5622.0000285. McNeill, R. L., B. E. Margason, and F. M. Babcock. 1965. “The role of soil
Choudhury, D., A. D. Katdare, and A. Pain. 2014. “New method to com- dynamics in the design of stable test pads.” In Proc., Guidance and
pute seismic active earth pressure on retaining wall considering seismic Control Conf., 366–375. New York, NY: American Institute of
waves.” Geotech. Geol. Eng. Int. J. 32 (2): 391–402. https://doi.org/10 Aeronautics and Astronautics.
.1007/s10706-013-9721-8. Murillo, C., L. Thorel, and B. Caicedo. 2009. “Ground vibration isolation
Davies, M. C. R. (1994). “Dynamic soil–structure interaction resulting with geofoam barriers: Centrifuge modeling.” Geotext. Geomembr.
from blast loading.” In Vol. 94 of Proc., Int. Conf. on Centrifuge 27 (6): 423–434. https://doi.org/10.1016/j.geotexmem.2009.03.006.

© ASCE 04018048-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048


Neumeuer, H. 1963. “Untersuchungen uber die Abschirmung eines Woods, R. D. 1968. “Screening of surface waves in soils.” J. Soil Mech.
bestehenden Gebaudes gegen Erschutterungen beim Bau and Betrieb Found. Eng. Div. 94 (4): 951–979.
einer U-Bahnstrecke.” [In German.] Baumaschine und Bautechnik 10 (1): Yang, Y. B., and H. H. Hung. 1997. “A parametric study of wave barriers
23–29. for reduction of train induced vibrations.” Int. J. Numer. Methods
Shahraki, M., M. R. S. Sadaghiani, K. J. Witt, and T. Meier. 2014. “3D Eng. 40 (20): 3729–3747. https://doi.org/10.1002/(SICI)1097-0207
modelling of train induced moving loads on an embankment.” Plaxis (19971030)40:20%3C3729::AID-NME236%3E3.0.CO;2-8.
Zakeri, J. A., M. Esmaeili, and S. A. Mosayebi. 2014. “Numerical inves-
Bull. 36 (Autumn): 10–15.
tigation of the effectiveness of a step-shaped trench in reducing
Wang, J. G., W. Sun, and S. Anand. 2009. “Numerical investigation on
train-induced vibrations.” Proc. Inst. Mech. Eng. Part F: J. Rail
active isolation of ground shock by soft porous layers.” J. Sound Vibr.
Rapid Transit 228 (3): 298–306. https://doi.org/10.1177/09544097
321 (3–5): 492–509. https://doi.org/10.1016/j.jsv.2008.09.047. 12473094.
Wang, Z. L., Y. C. Li, and J. G. Wang. 2006. “Numerical analysis of Zeng, X., J. G. Rose, and J. S. Rice. 2001. “Stiffness and damping ratio of
attenuation effect of EPS geofoam on stress-waves in civil defense en- rubber modified asphalt mixes: Potential vibration attenuation for high
gineering.” Geotext. Geomembr. 24 (5): 265–273. https://doi.org/10 speed railway track beds.” J. Vibr. Control 7 (4): 527–538. https://doi
Downloaded from ascelibrary.org by North Carolina A&T State Univ on 05/30/18. Copyright ASCE. For personal use only; all rights reserved.

.1016/j.geotexmem.2006.04.002. .org/10.1177/107754630100700403.

© ASCE 04018048-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2018, 144(8): 04018048

Vous aimerez peut-être aussi