You are on page 1of 111

REPUBLIC OF TURKEY

MUĞLA SITKI KOÇMAN UNIVERSITY


GRADUATE SCHOOL OF
NATURAL AND APPLIED SCIENCES

DEPARTMENT OF CIVIL ENGINEERING

INVESTIGATION OF VIBRATION ISOLATION


PERFORMANCE OF TRENCH TYPE
WAVE BARRIERS BY FIELD TESTS

MASTER OF SCIENCE

ONUR TOYGAR

DECEMBER 2015
MUĞLA
I hereby declare that all information in this document has been obtained and
presented in accordance with academic rules and ethical conduct. I also declare
that, as required by these rules and conduct, I have fully cited and referenced all
material and results that are not original to this work.

Onur Toygar
23/12/2015

iii
ABSTRACT
INVESTIGATION OF VIBRATION ISOLATION PERFORMANCE OF
TRENCH TYPE WAVE BARRIERS BY FIELD TESTS

Onur TOYGAR

Master of Science (M.Sc.)


Graduate School of Natural and Applied Sciences
Department of Civil Engineering
Supervisor: Assist. Prof. Dr. Deniz ÜLGEN
December 2015, 95 pages

Ground-borne vibrations caused by construction activities, highway and railway


traffic may disturb adjacent structures and dwellers. Thus, studies on isolation of this
type of vibrations have accelerated in recent years. Open and in-filled trench type
wave barriers have commonly been used in reducing unfavorable vibrations. Even
though there are great numbers of studies based on numerical methods about
vibration isolation performance of open and in-filled trenches, experimental data is
very limited. Therefore, vibration isolation performance of open and in-filled (EPS
geofoam and water) trenches were investigated by full scale field experiments in the
present study. Influence of vibration frequency, location of the vibration source, soil
layering, dimensions of the wave barriers and fill material type on screening
effectiveness was examined. First of all, an extensive site investigation was carried
out to determine physical and dynamic properties of the site. Later, field vibration
tests were performed in the absence of any wave barrier to observe the attenuation
characteristics of the site. Then, series of vibration tests were repeated for cases of
open trench, EPS geofoam wave barrier and water-filled trench. Isolation
performance of the wave barriers was determined quantitatively in frequency
domain. The results of the present study showed that screening effectiveness of the
wave barriers was mostly influenced by vibration frequency, Rayleigh wave length
of the soil, normalized trench depth and layering properties. It was concluded that the
depth of the barrier should have been at least Rayleigh wave length to achieve
reasonable isolation. Moreover, findings of the dissertation were compared with
those of reported in prior numerical and experimental studies and it was seen that
they were in a good agreement. Consequently open trench and EPS geofoam filled
trench outperformed water filled trench in isolation ground-borne vibrations.

Keywords: Vibration Isolation, Wave Barrier, EPS Geofoam, Wave Propagation,


Attenuation

iv
ÖZET
HENDEK TİPİ DALGA BARİYERLERİNİN TİTREŞİM YALITIMI
PERFORMANSININ SAHA DENEYLERİYLE İNCELENMESİ

Onur TOYGAR

Yüksek Lisans Tezi


Fen Bilimleri Enstitüsü
İnşaat Mühendisliği Anabilim Dalı
Danışman: Yrd. Doç Dr. Deniz ÜLGEN
Aralık 2015, 95 sayfa

İnşaat çalışmaları, otoyol ve demir yolu trafiği gibi kaynaklar nedeniyle oluşan yer
titreşimleri etrafındaki yapıları ve çevre sakinlerini rahatsız edebilir. Bu nedenle, bu
tip titreşimlerin yalıtımı ile ilgili çalışmalar son yıllarda hız kazanmıştır. İçi boş ve
dolu hendekler, istenmeyen titreşimlerin azaltılmasında dalga bariyeri olarak yaygın
bir şekilde kullanılmaktadır. İçi boş ve dolu hendeklerin titreşim yalıtım
performansını inceleyen, sayısal yöntemlere dayalı çok sayıda çalışma bulunmasına
rağmen deneysel veriler çok kısıtlı sayıdadır. Bu nedenle mevcut çalışmada içi boş
ve dolu (EPS ve su) hendeklerin titreşim yalıtım performansı tam ölçekli saha
deneyleriyle incelenmiştir. Titreşim frekansının, titreşim kaynağının yerinin, zemin
tabakalanmasının, dalga bariyerinin boyutlarının ve dolgu malzemesinin yalıtım
performansı üzerindeki etkisi incelenmiştir. Öncelikle zeminin fiziksel ve dinamik
özelliklerini belirlemek için kapsamlı bir zemin etüdü düzenlenmiştir. Daha sonra
zeminde titreşim azalım özelliklerini belirlemek için herhangi bir dalga bariyeri
yokken saha titreşim deneyleri yapılmıştır. Sonrasında içi boş, EPS dolu ve su dolu
hendek durumlarında saha titreşim deneyleri tekrarlanmıştır. Dalga bariyerlerinin
titreşim yalıtım performansı niceliksel olarak frekans uzayında bulunmuştur. Mevcut
çalışmanın sonuçları, dalga bariyerlerinin yalıtım performansının en çok titreşim
frekansından, zeminin Rayleigh dalga boyundan, normalize edilmiş hendek
derinliğinden ve zeminin tabakalanmasından etkilendiğini göstermiştir. Makul
derecede bir yalıtım elde etmek için hendek derinliğinin en az Rayleigh dalga boyu
kadar olması gerektiği sonucuna varılmıştır. Bunlara ek olarak, mevcut incelemenin
bulguları, önceki deneysel ve nümerik çalışmaların sonuçlarıyla karşılaştırılmış ve
birbirleriyle oldukça uyumlu olduğu görülmüştür. Sonuç olarak yer titreşimlerinin
yalıtılmasında içi boş ve EPS dolu hendekler, su dolu hendekten daha iyi bir
performans sergilemiştir.

Anahtar Kelimeler: Titreşim Yalıtımı, Dalga Bariyeri, EPS, Dalga Yayılımı,


Azalım

v
To my family,

vi
ACKNOWLEDGEMENTS

I would like to express my deepest and sincere gratitude to my supervisor


Assist.Prof.Dr. Deniz ÜLGEN for his technical guidance and encouragement
throughout the course of this dissertation. I would also like to thank him for his
kindly attitude, help and constant support.
I would also like to thank head of examining committee Assist.Prof.Dr. Mehmet
Rifat KAHYAOĞLU not only for his constructive comments about my thesis but
also his endless encouragement.
I am thankful to examining committee member Assist.Prof.Dr. Selman SAĞLAM for
his contributions and suggestions regarding the improvement of this dissertation.
I am thankful to Assist.Prof.Dr. Altuğ SAYGILI for teaching laboratory test and
helping me in this stage. His contributions are gratefully appreciated.
The study leading to these results has received funding from Muğla Sıtkı Koçman
University Scientific Research Projects with the BAP Code: 13-05. The financial
support from BAP Coordinatorship of Muğla Sıtkı Koçman University is gratefully
acknowledged.
EPS geofoam used in this study was supplied from ATERMİT İzmir factory. I am
thankful to Vise-President Hasan ÇİVİCİ for procurement of EPS.
I would also like to thank all of my colleagues at Civil Engineering Department and
employees of Engineering Faculty of Muğla Sıtkı Koçman University. Particularly, i
am indebted to İslam GÖKALP, Mesut DAĞDELEN, Mehmet KARATÜRK,
Hifzullah EFE and Doğukan DAYI for helping me in site experiments.
I wish to thank my close friends Oktay ERÖZKAN, Mehmet ŞAHİN and Kevser
KARABULUT for their constant support during the whole stages of my thesis.
Last but not the least, my deepest gratitude goes to my family and beloved girlfriend
Deniz for their unflagging love and support over years; this dissertation would be
simply impossible without them.

vii
TABLE OF CONTENTS

ACKNOWLEDGEMENTS ..................................................................................... vii


TABLE OF CONTENTS ........................................................................................ viii
LIST OF TABLES .................................................................................................... xi
LIST OF FIGURES ................................................................................................. xii
LIST OF SYMBOLS AND ABBREVATIONS ..................................................... xv
1. INTRODUCTION .................................................................................................. 1
1.1. General Overview .............................................................................................. 1
1.2. Objective of the Study ....................................................................................... 3
1.3. Outline of the Thesis ......................................................................................... 4
2. LITERATURE REVIEW...................................................................................... 6
2.1. Vibration Sources .............................................................................................. 6
2.2. Wave Propagation in Soil Medium ................................................................... 7
2.3. Wave Barriers .................................................................................................. 11
2.3.1. Open trenches ........................................................................................... 11
2.3.2. In-filled wave barriers............................................................................... 14
2.3.2.1. EPS geofoam wave barriers .............................................................. 14
2.3.2.2. Water-filled trenches ......................................................................... 16
2.3.2.3. Bentonite-filled trenches ................................................................... 17
2.3.2.4. Concrete wave barriers ..................................................................... 18
2.3.2.5. Gas cushions ..................................................................................... 20
2.3.2.6. Rubber chip wave barriers ................................................................ 20
2.3.2.7. Piles ................................................................................................... 21
2.3.2.8. Sheet piles.......................................................................................... 22
2.4. Effect of Barrier Geometry.............................................................................. 23
2.4.1. Source-barrier distance ............................................................................. 23
2.4.2. Width of barrier ........................................................................................ 25
2.4.3. Length of barrier ....................................................................................... 26
2.4.4. Depth of barrier ......................................................................................... 27
3. FIELD EXPERIMENTS ..................................................................................... 29
3.1. Site Investigation ............................................................................................. 29

viii
3.1.1. Physical properties .................................................................................... 30
3.1.2. Dynamic properties ................................................................................... 32
3.2. Test Equipment................................................................................................ 35
3.2.1. Vibration source........................................................................................ 35
3.2.2. Accelerometer ........................................................................................... 36
3.2.3. Data acquisition system ............................................................................ 37
3.3. Field Test Procedure ........................................................................................ 37
3.3.1. Open trench ............................................................................................... 39
3.3.2. In-filled trenches ....................................................................................... 41
3.3.2.1. EPS geofoam filled trench ................................................................. 42
3.3.2.2. Water filled trench............................................................................. 43
3.4. Calibration and Filtering of Data ..................................................................... 44
3.4.1. Calibration ................................................................................................ 44
3.4.2. Filtering of Data ........................................................................................ 44
4. RESULTS AND DISCUSSION .......................................................................... 46
4.1 Attenuation of Vibrations ................................................................................. 47
4.2. Vibration Isolation by Wave Barriers.............................................................. 51
4.2.1. Screening efficiency of open trenches ...................................................... 55
4.2.2. Screening efficiency of in-filled trenches ................................................. 58
4.2.2.1. Vibration isolation performance of EPS geofoam wave barriers ..... 58
4.2.2.2. Vibration isolation performance of water-filled trench .................... 60
4.3. Comparison on Screening Effectiveness of Wave Barriers ............................ 61
4.4. Comparison between Results of Current Study and Published Studies .......... 64
4.4.1. Experimental studies ................................................................................. 64
4.4.1.1. Open trench ....................................................................................... 64
4.4.1.2. EPS geofoam wave barrier ............................................................... 65
4.4.2. Numerical investigations .......................................................................... 67
5. SUMMARY AND CONCLUSIONS .................................................................. 69
5.1. Summary ......................................................................................................... 69
5.2. Conclusions ..................................................................................................... 69
5.3. Recommendations for Future Work ................................................................ 72
REFERENCES ......................................................................................................... 74
APPENDIX ............................................................................................................... 82
Appendix A. Results of Standard Penetration Tests .............................................. 82

ix
Appendix B. Results of Sieve Analyses ................................................................. 85
CURRICULUM VITAE .......................................................................................... 94

x
LIST OF TABLES

Table 3.1. Normalized dimensions of the barrier with respect to Rayleigh wave
length ........................................................................................................ 41
Table 4.1. Obtained amplitude reduction ratios in presence of wave barrier when
lt=2.5 m..................................................................................................... 62
Table 4.2. Obtained amplitude reduction ratios in presence of wave barrier when
lt=5.5 m..................................................................................................... 62
Table A.1. Standard penetration test log (Borehole 1)............................................... 82
Table A.2. Standard penetration test log (Borehole 2)............................................... 83
Table A.3. Standard penetration test log (Borehole 3)............................................... 83
Table A.4. Standard penetration test log (Borehole 4)............................................... 84
Table A.5. Standard penetration test log (Borehole 5)............................................... 84

xi
LIST OF FIGURES

Figure 2.1. Propagation of P-waves ............................................................................. 7


Figure 2.2. Propagation of S-waves ............................................................................. 8
Figure 2.3. Propagation of Rayleigh waves ................................................................. 9
Figure 2.4. Amplitude ratio of Rayleigh wave’s components vs. normalized depth by
Rayleigh wave length (Richart et al., 1970) ............................................ 10
Figure 2.5. Centrifuge test setup of Murillo et al. (2009) .......................................... 15
Figure 2.6. The schematic description of active isolation systems ............................ 23
Figure 2.7. The schematic description of passive isolation systems .......................... 24
Figure 3.1. Test site (a) location of Bayır (b) satellite view of the site...................... 29
Figure 3.2. Locations of field tests ............................................................................. 30
Figure 3.3. Corrected SPT blows (N60)’ record along boreholes ............................... 30
Figure 3.4. The soil profile of the site ........................................................................ 31
Figure 3.5. MASW test in the field (a) generation of impact (b) geophones ............. 32
Figure 3.6. Compression wave velocity profile of the site (a) north-south direction
(b) east-west direction ............................................................................. 33
Figure 3.7. Shear wave velocity profile of the site (a) north-south direction (b) east-
west direction (c) average ....................................................................... 34
Figure 3.8. Microtremor test (a) data acquisition in the field (b) SR04S3 triaxial
seismometer ............................................................................................. 35
Figure 3.9. Plate vibratory compactor ........................................................................ 35
Figure 3.10. MEMS type Sense Box 7001 accelerometer ......................................... 36
Figure 3.11. Test Box 2010 dynamic data logger ...................................................... 37
Figure 3.12. The location of accelerometers in the field after levelling .................... 38
Figure 3.13. Plan view of field test configurations (a) when source-barrier distance
(lt) was 2.5m (b) when source-barrier distance (lt) was 5.5m ................. 39
Figure 3.14. Illustration of dimensionless geometrical properties of the barrier ....... 40
Figure 3.15. Open trench in the field (a) general layout (b) closer view ................... 41
Figure 3.16. Determination of dynamic characteristics of EPS geofoam (a) Ultrasonic
Pulse Velocity Test (b) laboratory impact test ........................................ 42
Figure 3.15. EPS geofoam filled trench in the field ................................................... 43
Figure 3.18. Water filled trench in the field ............................................................... 44

xii
Figure 3.19. Application of 8-order Butterworth filter (a) f=25 Hz (b) f=50 Hz (c)
f=70 Hz.................................................................................................... 45
Figure 4.1. Obtaining peak particle acceleration amplitudes (in case of no trench
under vibrations having frequency of 25 Hz) from accelerograms (a) #1
(b) #2 (c) #3 (d) #4 (e) #5 (f) #6 .............................................................. 47
Figure 4.2. Variation of vibration attenuation when lt= 2.5m (a) no trench-open
trench (b) no trench-EPS geofoam filled trench (c) no trench-water filled
trench ....................................................................................................... 48
Figure 4.3. Variation of vibration attenuation when lt= 5.5m (a) no trench-open
trench (b) no trench-EPS geofoam filled trench (c) no trench-water filled
trench ....................................................................................................... 49
Figure 4.4. Obtaining Fourier Spectrums (in case of 4.5m deep open trench under
vibrations having frequency of 50 Hz) from accelerograms (a) #1 (b) #2
(c) #3 (d) #4 (e) #5 (f) #6 ........................................................................ 52
Figure 4.5. Attenuation of spectral amplitudes when lt= 2.5m (a) no trench-open
trench (b) no trench-EPS geofoam filled trench (c) no trench-water filled
trench ....................................................................................................... 53
Figure 4.6. Attenuation of spectral amplitudes when lt= 5.5m (a) no trench-open
trench (b) no trench-EPS geofoam filled trench (c) no trench-water filled
trench ....................................................................................................... 54
Figure 4.7. Variation of amplitude reduction ratio for open trench when (a) lt= 2.5m
(b) lt= 5.5m .............................................................................................. 56
Figure 4.8. Variation of amplitude reduction ratio for EPS geofoam wave barrier
trench when (a) lt= 2.5m (b) lt= 5.5m ...................................................... 58
Figure 4.9. Variation of amplitude reduction ratio for water filled trench when (a) lt=
2.5m (b) lt= 5.5m ..................................................................................... 60
Figure 4.10. Comparison of amplitude reduction ratios obtained from present study
with those of reported in experimental studies in the literature for case of
open trench .............................................................................................. 65
Figure 4.11. Comparison of amplitude reduction ratios obtained from present study
with those of reported in experimental studies in the literature for case of
EPS geofoam filled trench (a) lt=2.5 m (b) lt=5.5 m ............................... 66
Figure 4.12. Comparison of amplitude reduction ratios obtained from present study
with those of reported in numerical studies in the literature for case of
open trench .............................................................................................. 68
Figure B.1. Grain size distribution of the soil sample BH-1-15 (Borehole 1, depth
1.5m) ....................................................................................................... 85
Figure B.2. Grain size distribution of the soil sample BH-2-30 (Borehole 2, depth
3.0m) ....................................................................................................... 86
Figure B.3. Grain size distribution of the soil sample BH-1-60 (Borehole 1, depth
6.0m) ....................................................................................................... 87

xiii
Figure B.4. Grain size distribution of the soil sample BH-3-60 (Borehole 3, depth
6.0m) ....................................................................................................... 88
Figure B.5. Grain size distribution of the soil sample BH-1-75 (Borehole 1, depth
7.5m) ....................................................................................................... 89
Figure B.6. Grain size distribution of the soil sample BH-3-120 (Borehole 3, depth
12.0m) ..................................................................................................... 90
Figure B.7. Grain size distribution of the soil sample BH-1-135 (Borehole 3, depth
13.5m) ..................................................................................................... 91
Figure B.8. Grain size distribution of the soil sample BH-2-135 (Borehole 2, depth
13.5m) ..................................................................................................... 92
Figure B.9. Grain size distribution of the soil sample BH-1-225 (Borehole 1, depth
22.5m) ..................................................................................................... 93

xiv
LIST OF SYMBOLS AND ABBREVATIONS

α Ratio of Shear Wave Velocity to Compressional Wave Velocity


ρ Density of the Soil
Ɣ Unit Weight of the Soil
ν Poisson’s Ratio
λ Lame’s Constant
λr Rayleigh Wave Length
a Acceleration
Amax Maximum Spectral Amplitude
AR Amplitude Reduction Ratio
dt Trench Depth
D Normalized Trench Depth
E Young’s Modulus
f Frequency
G Shear Modulus
K Ratio of Rayleigh Wave Velocity to Shear Wave Velocity
lt Distance between Vibration Source and Wave Barrier
L Normalized Distance between Vibration Source and Wave Barrier
N60 Number of SPT Blow

N60 Corrected Number of SPT Blow
r Distance from Vibration Source
SV Vertical Component of Shear Wave Velocity
SH Horizontal Component of Shear Wave Velocity
T Period
VP Primary (Compressional) Wave Velocity
VR Rayleigh Wave Velocity
VS Shear Wave Velocity
wt Trench Width
W Normalized Trench Width
ADC Analog to Digital Converter
BEM Boundary Element Method

xv
BH Borehole
CH Clay of High Plasticity
CL Clay of Low Plasticity
EPS Expanded Polystyrene
FEM Finite Element Method
GWT Ground Water Table
LL Liquid Limit
MASW Multichannel Analysis of Surface Waves
MEMS Micro Electro-Mechanical System
PL Plastic Limit
SC Clayey Sand
SPT Standard Penetration Test
TRM Transmission and Reflection Matrix
UPS Uninterruptible Power Supply
USCS Unified Soil Classification System
XPS Extruded Polystyrene

xvi
1. INTRODUCTION

1.1. General Overview

Vibrations induced by frequent urban environment activities such as constructional


work, machinery actions, blasting, highway and railway traffic can cause detrimental
effects on adjacencies. These vibrations may even deteriorate the function of the
sensitive equipments within immediate vicinity as well as the life quality of the
dwellers. Thus, isolation of such undesired vibrations has become an important issue
in recent years. In order to control the effects of vibrations on nearby structures, soil
medium can be isolated by a suitable wave barrier. Open and in-filled trenches
(geofoam, concrete, water, bentonite, gas cushion, and rubber chip), sheet pile wall
and piles have been employed as wave barriers (Barkan, 1962; Woods, 1968; Al
Hussaini, 1992; Massarsch, 2005; Çelebi et al., 2009; Alzawi, 2011; Ju and Li, 2011,
Zoccali et al., 2015). It has been well addressed in the literature that type of vibration
source, operating frequency, physical and dynamic characteristics of the soil and
geometrical properties affect the performance of wave barriers.

Vibrations propagate in soil media as elastic waves which can be categorized under
two groups: body and surface waves. Miller and Pursey (1955) investigated the
scattering of vibrations generated by vertically oscillating footing located on the
surface. It was concluded that major part of the vibrations was surface waves such as
Rayleigh waves. Rayleigh waves attenuate at a slower rate than other elastic waves.
Since most of the vibration sources are located on the ground, Rayleigh wave is
primary concern of vibration isolation problems.

Vibration isolation cases were firstly reported by Barkan (1962) and McNeill et al.
(1965). The authors used open trench-sheet pile couple as wave barriers. Barkan
(1962) stated that isolation system could not reduce traffic induced vibrations.
McNeill et al. (1965) reported that sensitive laboratory equipment was satisfyingly
isolated by wave barrier couple.

1
In order to investigate isolation performance of wave barriers, some studies (Woods,
1968; Çelebi et al., 2009; Alzawi and El Naggar, 2011a) based on field experiments
were carried out. Woods (1968) examined effectiveness of open trench for cases of
active and passive isolation. Çelebi et al. (2009) compared screening efficiency of
open, bentonite and concrete filled trenches. Alzawi and El Naggar (2011a)
investigated isolation performance of open trench and geofoam wave barriers. The
authors concluded that open trench outperformed other barriers in the sense of
vibration isolation. It was also noted that efficiency of the barriers was mostly
influenced by trench depth.

Haupt (1981) conducted laboratory tests regarding open trench located on dense
sand. It was reported that amplitude of vibrations reduced by 50-85% in the presence
of open trench. Murillo et al. (2009) conducted centrifuge tests to evaluate isolation
performance of geofoam wave barriers. The authors concluded that dimension of the
wave barrier was vital with regard to screening effectiveness.

Parametric studies concerning vibration isolation performance of the wave barriers


have accelerated with development of numerical methodologies. Finite element
method (Haupt, 1978a; Haupt, 1978b; Wang et al., 2006; Ekanayake et al., 2014;
Saikia and Das, 2014; Zoccali et al., 2015) and boundary element method (Beskos et
al., 1986, Al Hussaini, 1992; Kattis et al., 1999a; Tsai and Chang, 2009) have been
commonly used to examine the isolation performance of wave barriers. Parameters
governing screening effectiveness were determined as vibration frequency and trench
geometry. Excitation frequency affects wave length of soil due to vibration. Since
major part of the vibrations propagates as Rayleigh wave, dimensions of wave barrier
were normalized with Rayleigh wave length in both experimental and numerical
studies. It was concluded that effectiveness of barriers was mostly influenced by the
normalized trench depth. Isolation performance can increase by increasing barrier
depth (Haupt, 1995; Al Hussaini and Ahmad, 1996; Tsai and Chang, 2009;
Ekanayake et al., 2014). The effect of normalized barrier width was examined and it
was concluded that width had negligible influence on screening efficiency (Beskos et
al., 1986; Ahmad and Al Hussaini, 1991, Wang et al., 2006; Lu and Tan, 2011;
Saikia and Das, 2014). Dasgupta et al. (1990), Ekanayake et al. (2014) and Zoccali et
al. (2015) investigated the effect of barrier length on isolation performance of open

2
and in-filled trenches. It was stated that efficiency of wave barrier was independent
from its length. Unlikely other numerical methods, Orehov et al. (2012) employed
finite difference method to study the behavior of open and in-filled trenches under
shock loading. The authors endorsed prior researches and concluded that screening
efficiency of the wave barriers was mostly influenced by barrier geometry. In order
to examine the soil-structure interaction in terms of train-induced vibrations,
coupling approach of finite element method and boundary element method was used
(Adam and von Estorff, 2005; Çelebi and Kırtel, 2013). It was reported that train-
induced vertical vibrations reduced in the presence of open and in-filled trench.

There are numerous studies regarding isolation performance of the wave barriers by
using numerical methods. In these studies, screening effectiveness of open trench and
in-filled barriers (geofoam, concrete, water bentonite, gas cushion, rubber chip, sheet
pile and piles) was investigated. Parameters affecting isolation performance of wave
barriers were determined. Inherently, assumptions such as elastic, isotropic and
homogeneous soil were incorporated in these numerical studies. Effect of soil
layering was not taken into consideration in most of the cases. Therefore, results
obtained from the numerical studies should be verified by field experiments
considering these parameters. However, number of case studies or studies based on
field tests is very limited. Increasing in situ data is to be a substantial mean to reveal
the vibration isolation performance of open trench and in-filled wave barriers.
Correspondingly, examining the influence of vibration frequency, trench dimensions
and location is to provide significant contribution on removing the inconsistencies
about isolation characteristics of the wave barriers.

1.2. Objective of the Study

This dissertation focuses on open and in-filled wave barrier performance under
continuous vibrations. The main objectives of the study can be summarized as
follows:

 To assess the vibration isolation performance of open and in-filled (EPS


geofoam and water) trenches under continuous vibration vertically generated
on the ground surface.
3
 To obtain the screening efficiencies of open and in-filled wave barriers in real
field conditions and compare them with those of published in the literature.

 To compare different wave barriers (open trench, EPS geofoam filled wave
barrier, water filled trench) in terms of vibration isolation and examine the
effect of fill material.

 To evaluate the effect of excitation frequency of vibration source.

 To investigate the influence of trench geometry on screening effectiveness of


the wave barriers.

In order to achieve listed goals, a series of full scale field experiments were
conducted. Vibrations with different frequencies were generated by a vibratory
compactor located different points on ground surface. A 4.5 m deep open trench was
excavated and filled with water and EPS geofoam separately. Amplitude of
vibrations was measured in vertical direction at specific points in absence and
presence of open trench, geofoam and water filled trenches. Efficiency of wave
barriers was determined quantitatively in different test configurations in real field
conditions. Finally, vibration isolation performance of wave barriers was compared
with those reported in prior studies.

1.3. Outline of the Thesis

This study is organized into five chapters. A brief summary of the thesis structure is
given as below:

 Chapter 1 gives brief information regarding vibration isolation of ground-


borne vibrations. The existing gaps and need for present study are described.
The objectives and scope of work are presented briefly.

 Chapter 2 provides a detailed literature review. Wave propagation mechanism


in soil medium is explained. Isolation performance of different wave barriers
and influence of governing parameters are discussed.

4
 Chapter 3 presents description of field test procedure. Properties of test
equipment are presented. Details of site investigation and field experiments
are explained.

 In Chapter 4, findings obtained from the present study are presented.


Screening effectiveness of the wave barriers (open and in-filled) at different
test configurations is discussed. Results of the present study are compared
with the findings of prior studies.

 Chapter 5 presents summary of the dissertation. Concluding remarks


regarding the vibration isolation performance of barriers are outlined.
Recommendations for further investigations are provided.

5
2. LITERATURE REVIEW

This chapter aims to give detailed information about vibration isolation. First, types
of vibration sources which cause unfavorable vibrations are presented. Since major
part of vibrations spread as surface waves, propagation of these waves is briefly
described. Then, types of wave barriers in reducing effects of vibrations are
thoroughly explained. Finally, influence of barriers’ geometrical properties is
discussed.

2.1. Vibration Sources

Vibration sources can be categorized as continuous or transient depending on type of


motion which they create (Anonymous, 2004). Continuous sources produce cyclic
motion such as vibratory pile driving or traffic, whereas transient vibration sources
generate single vibration events such as blasting or dropping impact hammer
(Brandenberg et al., 2010). Most of vibration isolation problems are arisen from
continuous sources. For example, construction activities or traffic loading generates
vibrations having peak velocity and frequency range of 1-50 mm/s and 10-60 Hz,
respectively (Henwood and Haramy, 2002). Meanwhile, human strongly perceives
vibrations having peak velocity more than 0.25 cm/s. Therefore, these vibrations
need to be reduced.

Woods (1968), Çelebi et al. (2009) and Alzawi (2011) utilized mechanical oscillators
to create vibrations in the field. Woods (1968) used a 33 kg vibration exciter to
produce vertical force of 18 pounds. Çelebi et al. (2009) utilized an electronic shaker
to generate harmonic load with amplitude of 250 N. Alzawi (2011) embedded a 79
kg oscillator 0.25 m below ground surface to produce sinusoidal force in a range of
4-60 Hz. On the other hand, Babu et al. (2011) used vibratory compactor to generate
vibrations having excitation frequency of 30-70 Hz. In brief, mechanical oscillators

6
and vibratory compactors have been commonly preferred as continuous vibration
sources in experimental studies.

2.2. Wave Propagation in Soil Medium

As aforementioned, there are lots of vibration sources disturbing the soil medium
such as pile driving, blasting, road and railway traffic and machinery foundations.
Since most of these sources are located at surface or near to ground, surface sources
are primary concern of vibration isolation. Vibrations originate from surface sources
propagate as elastic waves. These waves can be categorized into two types: body and
surface waves.

Body waves propagate through either pure volume change or pure rotation in soil
medium (Richart et al., 1970). The wave resulting from dilatation and compression in
soil is known as P-wave (primary wave) illustrated in Figure 2.1. The velocity of P
wave (Vp) is

  2G E(1  )
VP   (2.1)
 (1  )(1  2)

Where λ is the Lame’s constant, G is the shear modulus, ρ is the density of the soil, E
is the Young’s modulus and ν is the Poisson’s ratio. λ is defined as

E
 (2.2)
(1  )(1  2)

Figure 2.1. Propagation of P-waves

7
The type of body wave that originates from rotation without volume change is known
as S-wave (shear wave) illustrated in Figure 2.2. Shear wave is vector sum of two
perpendicular components such as SV (vertical) and SH (horizontal). The velocity of
S-wave (Vs) is

G
VS  (2.3)

Figure 2.2. Propagation of S-waves

When body waves reach the ground surface, surface waves occur by superposition of
body waves like a wave on a sea surface level as illustrated in Figure 2.3. This type
of surface wave is known as Rayleigh or Rayleigh-Lamb wave. The concept of
Rayleigh waves was firstly proposed by Rayleigh (1885) and later described by
Lamb (1904) while explaining boundary conditions of wave propagation. Since it is
originated from superposition of body waves, the velocity of Rayleigh wave (VR)
depends on both body waves’ velocity.

K 6  8K 4  (24  16 2 )K 2  16( 2  1)  0 (2.4)

Where K=VR/VS and α is

VS (1  2)
  (2.5)
VP 2(1  )

By rearranging (Eq. 2.4) and (Eq. 2.5), the relation between velocity of Rayleigh
wave and that of S-wave (Eq. 2.6) can be expressed as

8
0.86  1.14
VR  VS (2.6)
1 

Since Poisson’s ratio of soil varies between 0 and 0.5, velocity of Rayleigh wave
linearly changes from 0.86 to 0.95 VS.

Figure 2.3. Propagation of Rayleigh waves

As the waves radiate from the source, they gradually encounter larger volume of soil.
Thus, energy density in each wave decreases and this phenomenon also results in
attenuation of amplitude of motion. The decrease in the amplitude of the waves is
called as geometric (radiation) damping (Richart et al., 1970). Knowing r is the
distance from source, the amplitude of body waves decreases in soil medium and
along ground surface at a rate of 1/r and 1/r2, respectively. However, amplitude of
Rayleigh waves decreases in proportional to 1/r1/2. This means that geometrical
damping in Rayleigh waves is less than the body waves. Therefore, Rayleigh waves
possess more energy and amplitude at greater distances from vibration source than
those of body waves.

Miller and Pursey (1955) analyzed the dispersion of elastic waves originated from
vertically oscillating footing in homogenous, isotropic, elastic half space having
Poisson’s ratio of 0.25. Distribution of elastic waves transmitted to half-space was
reported as 67% Rayleigh wave, 26% S-wave and 7% P-wave. Sanchez-Sesma et al.
(2011) verified that most of the waves generated by vertically oscillating footing
spread as Rayleigh wave.

9
Richart et al. (1970) investigated the attenuation of Rayleigh wave’s amplitude with
soil depth. Half-space was modelled as homogenous, isotropic and elastic for
different Poisson’s ratios in the range of 0.25 and 0.5. Horizontal and vertical
components of Rayleigh wave were normalized with Rayleigh wave amplitude at
surface level. In this way, amplitude ratios for Rayleigh wave’s components were
derived as a function of normalized depth (Figure 2.4). As seen in Figure 2.4,
amplitude of Rayleigh wave decrease 75% at a normalized depth of 1.0λR, where λR
is the Rayleigh wave length. Woods (1968) suggested that an efficient wave barrier
should have achieved at least %75 reduction in vibration amplitudes. Thus, according
to Richart et al. (1970), the depth of a wave barrier should have been at least 1.0λR to
be regarded as successful in vibration isolation purpose.

Figure 2.4. Amplitude ratio of Rayleigh wave’s components vs. normalized depth by Rayleigh
wave length (Richart et al., 1970)

In brief, most of the vibration sources are situated at ground surface and surface
sources majorly generate Rayleigh waves. Since these waves attenuate at a slower
rate than body waves, attenuation of Rayleigh waves is a very important issue in

10
vibration isolation. Therefore, governing parameters such as the distance between
wave barrier and excitation source, the width, length, depth of wave barrier should be
taken into account in terms of Rayleigh wave length. Rayleigh wave length (λR) in a
soil medium due to a vertically oscillating vibration source can be calculated by
following equation (Eq. 2.7):

VR
R  (2.7)
f

Where f is operating frequency. In previous studies (Richart et al., 1970; Sanchez-


Sesma et al., 2011) half-space was assumed to be homogenous, isotropic and elastic.
However, some discrepancies can be encountered in real cases such as layering in
soil profile. Vibrations spread as elastic waves in soil medium. In a layered half
space, elastic waves may reflect or refract at layer interfaces. Thus, special attention
should be paid for propagation of vibrations, since isolation performance of wave
barriers depends on reflection, diffraction and scattering of Rayleigh waves (Jain and
Soni, 2007). Screening effectiveness of softer or stiffer barriers (than soil medium)
can change depending on the diffraction and reflection of Rayleigh waves around
barrier (Alzawi, 2011). As a result, wave propagation characteristics of soil medium
have great influence in vibration isolation problems.

2.3. Wave Barriers

2.3.1. Open trenches

The concept of using open trench as wave barrier was introduced by Barkan (1962).
The author employed open trench and sheet pile couple to reduce the effects of
traffic borne vibrations. However, the system turned out to be unsatisfactory (section
2.3.2.8). Woods (1968) conducted field experiments by using solitarily open trench.
A vertically oscillating vibration exciter was located on a circular footing as
vibration source. Circular trenches with different geometric configurations were
investigated to evaluate the vibration isolation performance of open trenches. In
order to measure the amount of isolation quantitatively, an amplitude reduction ratio

11
was defined by proportioning vibration amplitudes in presence and absence of
trench, respectively. As aforementioned, this ratio should be at most 0.25 for a
barrier to be deemed as efficacious. It was concluded that the trench depth should
have been at least 0.6λR and 1.19λR for active and passive isolation (section 2.4.1),
respectively. Çelebi et al. (2009) examined the screening effectiveness of open trench
in operating frequency range of 10-100 Hz. Due to the trench stability problem,
15cm concrete siding was utilized. Field test results were verified by using boundary
element method. The amplitude reduction ratio was obtained as 0.21. Since open
trench was only applicable for shallow depths by the virtue of stability issue, it was
reported that application of open trench was not practical. Alzawi and El-Naggar
(2011a) also conducted field experiments to examine the isolation performance of
open trenches and verified the results by 2D finite element analyses. The amplitude
reduction ratio was obtained in a range of 0.06-0.26 (averagely 0.16). The authors
endorsed Woods (1968) about trench depth should have been at least 0.6λR. It was
indicated that as the depth increased, higher reduction could have been achieved.

Haupt (1981) carried out laboratory tests on a 10m x 10m rectangular bin filled with
artificially densified sand. Measurements were performed 3.5λR after the trench.
Eventually, the amplitude reduction ratio was obtained with a range 0.15-0.5 in the
presence of open trench. Madheswaran et al. (2009) conducted small scale tests for
an open trench located in soft clay under pile driving conditions. Isolation
performance was measured in terms of peak particle acceleration. Ultimately,
vibrations due to pile driving were reduced around 75%. Then, the case was
modelled by using finite element code. It was reported that the FEM code
overestimated the peak particle acceleration by approximately 20%.

There are great number of researches based on boundary element method (Beskos et
al., 1986; Dasgupta et al., 1990; Ahmad and Al-Hussaini, 1991; Ahmad et al., 1996;
Kattis et al., 1999a; Tsai and Chang, 2009). Beskos et al. (1986) performed 2D
analyses to investigate screening effectiveness of open trenches and suggested that
the trench depth should have been at least 0.6λR as Woods (1968) stated. Ahmad and
Al Hussaini (1991) stated that isolation performance was influenced by geometrical
properties of the trench, particularly the depth of trench. Moreover the author
emphasized that width of the trench became more important in case of shallow trench

12
(when the depth was less than 0.8λR). Ahmad et al. (1996) developed a design
formulation in terms of geometrical properties of trench and soil properties and
validated the result of field tests conducted by Woods (1968). Dasgupta et al. (1990)
and Kattis et al. (1999a) investigated isolation performance of the open trenches by
using three dimensional boundary element analyses. Dasgupta et al. (1990) reported
that amplitude reduction was calculated as 0.29 and 0.312 for a line and surface,
respectively. Differently from Woods (1968) and Dasgupta et al. (1990), Kattis et al.
(1999a) suggested that the trench depth should have been at least 0.8λR. Tsai and
Chang (2009) investigated the effect of siding on open trench. Open trench with
0.5m concrete siding and 0.05m sheet pile siding was compared with purely open
trench. It was indicated that installation of any siding decreased the performance of
the trench. Moreover, as the depth of trench increased, higher level of vibration
isolation was obtained.

Jesmani et al. (2008) performed 3D finite element analyses for isolation of deep
foundations by open trench. It was recommended that trench depth should have been
at least half length of the pile. Wang et al. (2009) carried out 3D analyses by using
LS-DYNA finite element program to examine blast induced vibrations on buried
structures. Open trench showed better performance when compared with filled
barriers (geofoam, concrete and water filled trenches). However, open trench was not
convenient as permanent measure since it could have been collapse under blasting.
Lu and Tan (2011) investigated the building response under deep dynamic
compaction impacts. Vibrations were reduced by 90% in the presence of open trench.
As the depth of trench increased, settlement and acceleration on building decreased.
Çelebi and Kırtel (2013) employed finite element analyses to observe the soil-
structure interaction under train-induced vibrations. Vertical vibrations were reduced
by 85% in case of open trench.

Coupling approach of the finite element method and boundary element method was
employed to investigate the train-induced vibrations (Adam, 2002; Adam and von
Estorff, 2005; Andersen and Nielsen, 2005). It was concluded that isolation
performance of open trenches mainly depended on depth of barrier. Deep open
trench was the best solution among other barriers. Inasmuch as depth of trench
increased, better isolation performance could have been achieved.

13
Unlike other numerical methods, Orehov et al. (2012) performed 2D finite difference
analyses by using FLAC program. Screening effectiveness of open trench under
shock loading (maximum load of 1000kN) was examined. Trench depth varying
from 5m to 20m was employed. Soil particle displacement decreased as the depth
increased. Consequently, vibration isolation performance of open trenches essentially
depends on dimensions of barrier. The influence of geometrical properties of barriers
is explained at section 2.4. On the other hand, it has obviously seen that there are vast
number of numerical studies but researches based on field experiments are very
limited. Thus, further field experiments are needed to understand isolation
mechanism of open trenches and to verify numerical studies.

2.3.2. In-filled wave barriers

2.3.2.1. EPS geofoam wave barriers

Geofoam term was defined as all types of plastic foams used for any geotechnical
purposes (Horvath, 1992). There are two types of geofoams according to their
production processes such as EPS (expanded polystyrene) and XPS (extruded
polystyrene). Since this material is a new member of geosynthetics family, there are
only a few studies examining the dynamic properties of geofoam. Athanasopoulos et
al. (1999) conducted resonant column tests and cyclic uniaxial tests to determine the
dynamic properties of EPS geofoam. It was concluded that EPS geofoam behaved as
an elastic material and had very low damping ratio for strain amplitudes up to 0.1%.
It behaves nonlinearly and has higher damping ratios within the range of 0.1% to
10%. Erickson (2011) employed cyclic triaxial tests to understand the dynamic
behavior of EPS geofoam. It was observed that damping ratio of EPS geofoam
increased with increase of confining pressure. Padade and Mandal (2012) indicated
that shear strength of EPS geofoam was directly proportional to its density.

Among geofoam types, EPS is the most commonly used one in civil engineering
applications (Horvath, 1996). Amini (2013) investigated seismic stability of EPS
embankments. It was observed that EPS geofoam embankments could resist dynamic
motion of 0.6g without interlayer sliding. Kafash (2013) examined landslides
stabilized with EPS. A seismic analysis procedure for installation of EPS blocks in
14
landslide stabilization was recommended. Wang et al. (2006) and Qiu (2015) used
finite element method to observe effects of EPS geofoam under explosion. Wang et
al. (2006) proposed an empirical formula for calculation of EPS inclusion length in
civil defense structures. Qiu (2015) developed an optimization design method in use
of EPS geofoam for instantaneous and time-delayed explosions for structures such as
tunnels. Alternatively, EPS can be used as a fill material in wave barriers. Alzawi
and El Naggar (2011a) conducted a series of full scale field tests to investigate the
screening effectiveness of EPS-filled trenches. A 3.0m deep thin trench was
constructed by hydro-dig method and filled with geofoam. Filled trench was exposed
to vertical harmonic loading within excitation frequency range of 40-60 Hz.
Consequently, the authors reported that EPS barriers could provide 68% reduction in
vibration amplitude. Moreover, it was indicated that more reduction could have been
achieved as the depth of the geofoam barrier increased. Murillo et al. (2009)
employed centrifuge tests (Figure 2.5) to examine the effect of EPS wave barriers on
isolation of vibrations generated by traffic loading. It was stated that geofoam
barrier’s dimensions were vital in the sense of vibration isolation. Accordingly, the
barrier was unfavorable when it was located less than 0.5 λR or more than 1.2 λR
away from vibration source. Results showed that EPS geofoam barrier was
efficacious when barrier’s width and depth were greater than 0.25λR and 1.5λR,
respectively.

Figure 2.5. Centrifuge test setup of Murillo et al. (2009)

There are also some numerical studies regarding EPS wave barriers’ vibration
screening performance. Wang et al. (2009) used finite element method for analyzing
the performance of EPS geofoam under ground shock. It was concluded that EPS
geofoam could be used as a wave barrier material in practice. Alzawi and El Naggar
(2009, 2011b), Qiu et al. (2014) and Ekanayake et al. (2014) performed 2D/3D
15
numerical analyses to investigate the geofoam wave barriers’ screening effectiveness
using ABAQUS finite element program. Alzawi and El Naggar (2009) reported that
screening effectiveness of all geofoam barriers fluctuated between 38% and 80%.
Alzawi and El Naggar (2011) and Qiu et al. (2014) stated that Poisson ratio and
material damping had negligible effect in vibration reduction. Ekanayake et al.
(2014) observed that efficiency of geofoam barrier could increase by depth. As a
result, it was concluded that EPS geofoam was beneficial as a fill material in wave
barriers.

2.3.2.2. Water-filled trenches

Due to the structural instability of open trenches, using fill material is a convenient
approach to overcome this difficulty. For this reason, water can be used in filling the
open trench barriers. Unfortunately, there are very limited studies concerning the
vibration isolation performance of water-filled trenches. Çelebi et al. (2009)
conducted field experiments to observe the effects of water-filled trench. Trench
consisting of concrete side walls was filled with water and isolation performance was
examined for different frequencies of harmonic loading. The authors concluded that
water-filled trench was efficient at only low frequencies. Wang et al. (2009) and Ju
and Li (2011) performed 3D numerical analyses for investigating the behavior of
water-filled trenches by using finite element method. Wang et al. (2009) reported
that inundated water trench had almost no effect in reduction of blast induced ground
shock. Ju and Li (2011) investigated isolation performance of the open trenches filled
with water at different heights. Water barriers enabled same effectiveness as open
barriers in attenuation of vertical waves, but passage of horizontal compression
waves could lead to underperformance of these barriers. The authors explained that
the incident waves and the waves coming from bottom of trench were cancelled out
each other in vertical direction. However, this phenomena was not valid for
horizontal direction since shear waves could not travel through the water. Orehov et
al. (2012) used FLAC finite difference program to examine the screening
effectiveness of water barriers. In the study, water trench with concrete walls was
subjected to shock loading. Results showed that soil particle displacement was
reduced owing to the presence of water barrier. Findings in the literature can not give

16
certain comments regarding the vibration isolation performance of water-filled
trenches. Consequently, further investigations are needed to conclude a general
opinion for utilization of water as fill material in wave barriers.

2.3.2.3. Bentonite-filled trenches

Bentonite can be considered as a soft wave barrier for vibration isolation purpose
(Çelebi et al., 2009). Çelebi et al. (2009) conducted field experiments to study the
effects of bentonite-filled trenches on screening effectiveness. A concrete walled-
trench was filled with bentonite and tested for both active and passive isolation cases.
Results showed that bentonite was more effective than other fill materials water and
concrete. Al Hussaini and Ahmad (1996) employed boundary element method to
determine the isolation performance of bentonite wave barriers. It was concluded that
the depth of the barrier should be greater than 0.8λR. El Naggar and Chehab (2005),
Lu and Tan (2011) and Zoccali et al. (2015) used finite element method to study the
behavior of bentonite-filled trenches. El Naggar and Chehab (2005) investigated
bentonite wave barriers’ performance under impact loading in time domain. It was
observed that bentonite-filled trenches outperformed than stiff barriers. Lu and Tan
(2011) studied building responses to deep dynamic compaction impact in terms of
acceleration and reported that a reduction of 90% can be achieved in presence of
bentonite-filled trenches. Zoccali et al. (2015) investigated filled trenches’ vibration
mitigation capacity under train induced ground vibrations and stated that 20%
reduction in vibration isolation can be obtained. Adam and von Estorff (2005)
investigated soil-structure interaction in terms of train induced vibrations in time
domain. A six story building and soil were modelled by using boundary and finite
element method couple. It was reported that vertical vibrations in building were
reduced to a level of 80%. All numerical studies showed that soft barriers were more
efficient than stiff barriers. Moreover, as the depth of bentonite wave barrier
increased, vibration isolation performance of the barrier increased as well.

17
2.3.2.4. Concrete wave barriers

Concrete wave barrier can be considered as a stiff wave barrier due to its higher
density compared to that of soil. Haupt (1981) performed small-scale laboratory tests
on artificially densified sand to examine screening effectiveness of the concrete wave
barriers. The results showed that amplitude of vibrations was reduced by 50%.
Çelebi et al. (2009) conducted full scale field experiments to compare the isolation
performance of open trenches and concrete wave barriers. Open trenches were
constructed with 15cm concrete thick walls to keep the stability of the trenches. The
findings indicated that the vibrations could be reduced by around 35% using concrete
wave barriers and open trenches showed a better isolation performance than
concrete-filled trenches.

Apart from experimental researches, there are a great number of numerical studies
concentrated on concrete wave barriers. Haupt (1977; 1978a; 1978b), El Naggar and
Chehab (2005), Madheswaran et al. (2009), Lu and Tan (2011) and Çelebi and Kırtel
(2013) used finite element method to investigate concrete wave barriers in 2D. Haupt
(1977; 1978a; 1978b) examined different geometric configurations of rectangular
concrete walls. It was reported that vibration isolation performance of concrete
barriers mainly depended on the cross sectional area of wall. As the cross sectional
area of barrier increased, amplitude reduction increased as well. Moreover, it was
emphasized that depth of trench had a greater effect on vibration isolation than the
width of trench. El Naggar and Chehab (2005) indicated that softer barriers
outperformed than concrete-filled trenches. Madheswaran et al. (2009) concluded
that effectiveness of concrete-filled trenches depended on the strength of the
concrete. Lu and Tan (2011) and Çelebi and Kirtel (2013) investigated building
response to ground vibrations. Lu and Tan (2011) stated that concrete barriers had
negligible effect in mitigation of acceleration due to vibration. Çelebi and Kırtel
(2013) reported 50% reduction in acceleration for short period buildings (T=0.25-0.5
s). Wang et al. (2009) and Zoccali et al. (2015) performed 3D finite element analyses
to investigate the vibration isolation performance of concrete filled trenches. Wang et
al. (2009) indicated that effect of concrete wave barrier on isolation of ground shock
was limited. Zoccali et al. (2015) reported that the train induced ground vibration
could only be reduced by 20%.
18
Boundary element method (BEM) is more suitable for wave propagation problems
(Haupt, 1995). Hereat, Beskos et al. (1986), Ahmad and Al-Hussaini (1991) and Tsai
and Chang (2009) employed BEM algorithm to comprehend 2D vibration isolation
issues. Beskos et al. (1986) examined isolation performance of trenches with
numerical examples. The authors concluded that concrete wave barriers’ cross
sectional area should have been greater than 1.5λR2. Ahmad and Al-Hussaini (1991)
noted that most important parameters of concrete-filled trenches were the
geometrical dimensions of barriers. Furthermore, the authors stated that the shear
modulus and the density of fill material had also effect on vibration isolation and
Poisson’s ratio could be ignored. Results of the study showed that an amplitude
reduction of 70% was obtained. Tsai and Chang (2009) investigated the influence of
concrete siding in vibration isolation performance of open trenches. It was indicated
that open trenches outperformed than those with concrete siding. In addition,
concrete diaphragm walls as wave barriers were only effective for shallow depth
(d<1.5λR). Dasgupta et al. (1990), Al Hussaini and Ahmad (1996) and Kattis et al.
(1999a) carried out 3D boundary element analyses to examine the vibration isolation
performance of concrete wave barriers. Dasgupta et al. (1990) calculated the
amplitude reduction ratio over a specific area rather than a certain point or a line.
Thereby, the amplitude reduction ratio after trench was calculated as 0.258. Al
Hussaini and Ahmad (1996) stated that the distance between vibration source and
concrete-filled trench should have been in the range of 0.2λR and 0.375λR. As the
depth and the width of trench increased, screening effectiveness increased, as well.
However, effect of width remained same for trenches wider than 0.5λR. Kattis et al.
(1999a) compared concrete trenches and piles with regards to vibration isolation.
Trenches provided better performance than piles. Depth of barrier should be greater
than 0.8λR for either trenches or piles.

Andersen and Nielsen (2005) employed FEM and BEM couple to analyze vertical
and horizontal vibrations along railway. No obvious solution was achieved in
horizontal direction. At low frequencies, concrete-filled trenches performed better
than soft barriers vertically. Their performance was same at high frequencies. Orehov
et al. (2012) used finite difference method in examining the building response under
shock loading. Open trench and water filled trench with concrete casing and concrete
wave barrier were compared in terms of acceleration reduction when the trench depth
19
changed between 5m to 20m. It was concluded that all types of wave barriers
performed well. Consequently, concrete filled trenches can be used as stiff wave
barriers.

2.3.2.5. Gas cushions

Low-density materials can be used as soft wave barrier to achieve a reasonable


isolation efficiency. When they are subjected to high lateral earth pressure, these
materials become stiffer and lose their isolation characteristic (Massarsch, 2005). In
this case, gas-filled cushions can be proposed to screen the ground-borne vibrations.
It is covered with a protective layer formed from polymeric thin film and aluminum
which is impermeable to gas. Gas cushions are inflated as same pressure as
surrounding earth pressure and utilized as fill material in trenches. Massarsch (2005)
investigated isolation performance of gas cushions by using the boundary element
method. Findings of the study showed that it had approximately the same
effectiveness as open trench. The author also conducted full-scale field experiments
on deep soft clay deposit in Uppsala. Gas cushions were installed in an open trench.
Then, it was filled with a cement-bentonite slurry to protect the cushions against
puncturing permanently. Results of the theoretical study and field experiment were in
good agreement. It was concluded that gas-filled cushion barriers with a depth of at
least 1.0λR could provide vibration isolation screening of 50-80%. El Naggar and
Chehab (2005) examined behavior of gas cushions under impact loading by using the
ANSYS finite element program. The authors stated that gas-filled cushions were
more effective than stiff barriers and isolation performance could increase directly
proportional to the depth of barrier.

2.3.2.6. Rubber chip wave barriers

Rubber chip generally consists of slivered waste tires. It has high damping ratio
(10%) and lower density (ρ=500kg/m3) than the soil matrix (Andersen and Nielsen,
2005; Buonsanti et al., 2009). Therefore, it is a good alternative to be used as a wave
barrier. Although there is no experimental study concerning isolation performance of
the rubber chips, some numerical studies (Andersen and Nielsen, 2005; Buonsanti et

20
al., 2009; Zoccali et al., 2015) employed rubber chips to reduce ground vibrations
induced by trains. Andersen and Nielsen (2005) used finite element-boundary
element method couple in frequency domain to examine the isolation of train-
induced ground vibrations. They stated that no regular trend was observed regarding
screening effectiveness of the rubber chips. Amplitude reduction ratio irregularly
changed with the excitation frequency. As the speed of the train increased, efficiency
of the barrier increased to 50%. Buonsanti et al. (2009) used ADINA finite element
program to model the rubber chip barrier in elastic half space. The authors concluded
that high damping ability of rubber chip didn’t promote its isolation performance
reasonably. Zoccali et al. (2015) performed 3D finite element analyses to study the
effect of rubber chip filled wave barrier on vibration isolation. However, only 5%
reduction in peak particle velocity was reported. As a result, numerical investigations
have not shown substantial isolation effect of rubber chip barriers, up to the present.

2.3.2.7. Piles

Utilization of piles as wave barriers is a new field in vibration isolation. Therefore,


no experimental study about screening efficiency of the piles was encountered.
Firstly, Aviles and Sanchez-Sesma (1983; 1988) numerically investigated screening
effectiveness of piles in both 2D and 3D space. Rigid circular piles were examined as
barriers for elastic waves. It was concluded that length and diameter of pile should
have been greater than 2λR and 0.25λR to achieve 50% efficiency, respectively. Kattis
et al. (1999a; 1999b) used 3D boundary element analyses in frequency domain.
Tubular and concrete solid piles were compared with equivalent open and concrete
trenches, respectively (Kattis et al., 1999a). Trenches always outperformed piles and
tubular hollow piles had best screening efficiency among piles. Amplitude reduction
was reported in a range of 0.3 and 0.6 for all piles. Kattis et al. (1999b) employed
row of piles having different cross-sectional shapes such as tubular, solid, circular
and square. The authors concluded that shape of pile did not have noteworthy
contribution to isolation performance. The spacing between piles should have been
small for reasonable vibration isolation. Gao et al. (2006) modelled layout of Kattis
et al. (1999b) and increased number of rows by using integral equations to govern
wave scattering. It was indicated that amplitude reduction ratio of 33% was obtained

21
for three rows of piles. Tsai et al. (2008) studied the effect of material in piles by
using FORTRAN 3D boundary element program. Steel pipes, concrete hollow and
solid piles and timber piles were examined in the sense of isolation performance.
Steel pipes showed the best performance having amplitude reduction ratio less than
0.2. Concrete solid and timber piles outperformed concrete hollow piles by
possessing amplitude reduction ratio of 0.4. Moreover, it was remarked that the most
important parameter was pile length in all cases. Lu et al. (2009) verified the findings
of Kattis et al. (1999b) by using TRM (Transmission and Reflection Matrix) method
and utilized same layout in layered medium against moving load induced vibrations.
Better efficiency was obtained for higher speeds of source when upper layer was
softer. Xu et al. (2010) numerically studied the influence of piles’ elasticity and
stated that stiff piles outperformed flexible ones. It was concluded that long piles
with small spacing should have been used for acquiring better screening
effectiveness (Lu et al., 2009; Xu et al., 2010; Lu et al., 2013). Turan et al. (2013)
performed 3D finite element analyses to observe the behavior of inclined secant
micro-piles in active isolation. Effectiveness mainly depended on inclination angle
and length of piles and was independent from distance between source and barrier.
As inclination angle was increased from 75° to 120°, amplitude reduction was
obtained in a range of 20% to 69%, respectively. Consequently, it was clearly
understood that vibration isolation performance of piles depended on length and
spacing of piles. Nevertheless, further studies are needed due to lack of experimental
investigations to verify numerical findings.

2.3.2.8. Sheet piles

Installation of sheet piles on the purpose of vibration isolation was firstly endeavored
by Barkan (1962). A building was desired to be isolated from traffic induced
vibrations by using open trench and sheet pile couple. However, unfavorable
vibrations had still remained and it appeared that the system was inefficacious. The
author stated that dimensions of steel sheets should have been much bigger. McNeill
et al. (1965) used a similar system consisted of open trench having width of 90 cm
and sheet pile having length of 9 m in protection of laboratory having sensitive
equipment from unfavorable vibrations. It was reported that the laboratory was

22
satisfyingly isolated. Woods (1968) conducted field tests to compare screening
effectiveness of open trenches and aluminum sheet piles in passive isolation
conditions. As a result, open trench outperformed sheet pile wall in reducing ground
borne vibrations. Apart from field experiment, some numerical analyses were also
performed. Tsai and Chang (2009) employed boundary element method to observe
the effect of siding on open trenches. It was concluded that sheet pile with open
trench was more effective in deep depths (d>1.5λR) and 90% reduction was obtained
in this case. Lu and Tan (2011) employed finite element analyses to investigate a
three-storey structure’s response against deep dynamic compaction impacts in the
presence of sheet pile. No significant effect was reported in reducing acceleration
due to the impact loading. Nevertheless, the authors stated that influence of depth
was more important than that of width for stiff barriers.

2.4. Effect of Barrier Geometry

2.4.1. Source-barrier distance

Vibration isolation can be divided into two groups with respect to source-barrier
distance: active and passive isolation (Woods, 1968). When the wave barriers are
located near the vibration source, it is known as active isolation as shown in Figure
2.6. On the contrary, in passive isolation, the wave barriers are installed near the
structure or site desired to be protected from unwanted vibrations (i.e. far from
vibration source) as shown in Figure 2.7.

Figure 2.6. The schematic description of active isolation systems

23
Figure 2.7. The schematic description of passive isolation systems

Woods (1968) conducted field experiments by using open trench for both active and
passive isolation conditions. Isolation performance of the trench depended on its
depth. Different barrier depths were suggested to achieve reasonable isolation for
active and passive isolation cases (see chapter 2.4.4). Çelebi et al. (2009) performed
field tests to compare isolation types. It was stated that passive isolation was more
effective than active isolation. Alzawi and El Naggar (2011a) investigated the
influence of vibration source-wave barrier distance on the efficiency of open and
EPS geofoam filled trenches by field experiments. Results of those experiments
showed that, deeper trench was required as the distance between source and barrier
increased.

Murillo et al. (2009) conducted centrifuge tests using geofoam wave barriers. The
barrier was insufficient when located less than 0.5λR away from the vibration source.
Depending upon the attenuation level, the geofoam wave barrier was also
inefficacious at greater distances. Thus, it was concluded that the wave barrier should
have been located at a distance within a range of 0.5λR and 1.2λR away from
vibration source.

Ahmad and Al Hussaini (1991) employed boundary element method to understand


the behavior of open and concrete trenches in passive isolation case. The trench was
located in a range of 3λR -12λR far from the source. It was seen that the distance
between source and barrier had negligible effect in passive isolation. Ahmad et al.
(1996) and Al Hussaini and Ahmad (1996) examined active isolation of machine
foundations by using a BEM algorithm. It was concluded that isolation performance
of the barrier highly depended on the source-barrier distance in active case. In

24
addition, Al Hussaini and Ahmad (1996) suggested that this distance should have
been 0.375λR.

Hamissi and Vileh (2007) performed 3D finite element analyses to study the effect of
source barrier distance on isolation performance of open trenches. It was reported
that an amplitude magnification could ingenerate due to reflection of waves in front
of the trench. Thus, the source-barrier distance was important aspect of unprotected
area (in front of the trench). Saikia (2014) examined the behavior of a pair of open
trench using the finite element method. The author concluded that the trench location
has no influence on screening effectiveness. Alzawi and El Naggar (2009) and Qiu et
al. (2014) employed finite element analyses to analyze the effect of trench-barrier
distance on isolation performance of EPS filled trenches. Alzawi and El Naggar
(2009) emphasized that screening effectiveness of a wave barrier was not affected by
location, whereas dimensions of the barrier have more influence on the vibration
isolation efficiency. Qiu et al. (2014) indicated that trench location did not have
remarkable contribution, nevertheless the isolation system should have been installed
near to the vibration source. Ekanayake et al. (2014) carried out 3D finite element
analyses by water-filled trenches. The authors also confirmed that the trench location
did not have significant effect on screening effectiveness.

2.4.2. Width of barrier

Influence of wave barriers’ width on vibration isolation has been investigated by


both field experiments and numerical studies. Woods (1968) employed field tests and
revealed that the width was not important with regards to isolation performance of
wave barriers. Murillo et al. (2009) conducted centrifuge test series to examine the
screening effectiveness of geofoam wave barriers. The effect of width was not
noteworthy for deep wave barriers, whereas it had considerable influence for shallow
depths.

Beskos et al. (1986), Ahmad and Al Hussaini (1991), Saikia and Das (2014)
performed numerical analyses to analyze the behavior of open trenches. It was
concluded that influence of the trench depth was ignorable. Ahmad and Al Hussaini
(1991) endorsed the results of Murillo et al. (2009) and stated that effect of barrier

25
became important for only shallow depths. Shrivastava and Kameswara Rao (2002)
and Lu and Tan (2011) investigated the behavior of open trench under impact
loading. The authors concluded that trench width had no observable effect on
attenuation of ground acceleration. Adam and von Estorff (2005) and Ekanayake et
al. (2014) promoted the importance of the trench depth. Increasing the depth of
barrier was suggested instead of widening the wave barrier.

Some numerical studies have investigated the effect of width on screening


effectiveness of in-filled trenches. Al Hussaini and Ahmad (1996) indicated that the
effect of concrete wave barriers’ width was considerable when it was smaller than
0.5λR. The effect diminished beyond this level. On the contrary to open trenches,
Beskos et al. (1986) stated that the width strongly affected the isolation performance
of concrete barriers. Wang et al. (2006) studied EPS geofoam inclusions under
blasting. It was observed that the effect of geofoam barrier’s width was barely
noticeable. It may be concluded that the width of wave barrier has negligible effect
on screening effectiveness.

2.4.3. Length of barrier

Dasgupta et al. (1990) performed 3D boundary element analyses with open trenches.
Two identical, alongside trenches (each) having length of 0.4λR were compared with
an another trench having length of 1.3λR. It was concluded that one big trench
outperformed in terms of vibration isolation. Shrivastava and Kamesara Rao (2002)
examined open trenches under impulse loading. Better isolation was achieved by
longer trenches. Ekanayake et al. (2014) indicated that the length of the barrier had
no distinct effect on screening effectiveness for both geofoam and water filled
trenches. Zoccali et al. (2015) employed concrete, rubber chip and bentonite filled
trenches having length range of 4m to 50m. It was seen that the influence of the
barriers’ length was not appreciable. As a result, it may be concluded that vibration
isolation performance of a wave barrier is independent of its length.

26
2.4.4. Depth of barrier

There are numerous studies concerning with governing parameters in screening


effectiveness of wave barriers. It was concluded that isolation performance was
strongly dependent on the depth of the barrier (Murillo et al., 2009; Tsai and Chang,
2009; Di Mino et al., 2009; Jesmani et al., 2011). Woods (1968) carried out field
tests to investigate screening effectiveness of open trench as wave barrier. It was
suggested that the trench depth should have been greater than 0.6λR and 1.19λR, for
active and passive isolation, respectively. Alzawi and El Naggar (2011a) indicated
that isolation performance of both open and EPS geofoam filled trenches was
essentially influenced by the depth. According to field test results, the depth of both
barriers should be greater than 0.6λR.

Murillo et al. (2009) revealed that isolation performance of geofoam wave barrier
was a function of barrier’s depth. Centrifuge test results showed that the barrier’s
depth should have been greater than 1.5λR to achieve reasonable isolation. As the
barrier’s depth increased, higher reduction could have been obtained.

Dasgupta et al. (1990) used boundary element method to understand the 3D nature of
open trenches’ isolation characteristics. It was stated that trench depth should have
been at least 0.6λR. Al Hussaini and Ahmad (1996) recommended that depth of
bentonite filled trench should have been greater 0.8λR. Kattis et al. (1999a) compared
screening effectiveness of open trenches and pile barriers. Even though open
trenches outperformed than piles, it was concluded the depth of both barriers should
have been greater than 0.8λR. Massarsch (2005) used gas-filled cushion barriers for
vibration isolation purpose. It was reported that 50-80% of screening effectiveness
could have been obtained by 1.0λR deep barriers. Jesmani et al. (2008) investigated
active isolation of deep foundations by using finite element method. The authors
expressed that trench should have been deeper than at least half length of the piles.
Alzawi and El Naggar (2011b) suggested that depth of 1.2λR could have been used to
obtain maximum efficiency in practical design of EPS wave barriers. There have
been many comments regarding the minimum depth of barrier to achieve sufficient
isolation. Independent from minimum depth, Haupt (1995), Al Hussaini and Ahmad
(1996), Shrivastava and Kameswara Rao (2002), Adam and von Estorff (2005), El

27
Naggar and Chehab (2005), Tsai and Chang (2009), Lu and Tan (2011), Jesmani et
al. (2011) and Ekanayake et al. (2014) concluded that better isolation could have
been obtained as the depth of the barrier increased.

28
3. FIELD EXPERIMENTS

3.1. Site Investigation

Field experiments were carried out in a flat area located in Bayır (Muğla, Turkey).
The site (Figure 3.1) was far from any vibration sources such as traffic, construction
actives etc. Thus, it was feasible to conduct vibration tests in this area. Prior to
vibration tests, a detailed site investigation was performed to evaluate the physical
and dynamic characteristics of the soil.

Figure 3.1. Test site (a) location of Bayır (b) satellite view of the site

29
3.1.1. Physical properties

In order to determine physical properties of the soil, series of in-situ and laboratory
tests were carried out. First of all, locations of boreholes were determined as points
where were close to the center and the corners of the site as shown in Figure 3.2.

Figure 3.2. Locations of field tests

Five boreholes were drilled in a depth range of 10m to 30m and soil samples were
continually obtained. Therewithal, Standard Penetration Tests (SPT) were conducted

Figure 3.3. Corrected SPT blows (N60)’ record along boreholes


30
at every 1.5m in boreholes. The number of SPT blows (N60) were recorded and
required corrections were carried out. The variation of corrected SPT blows (N60)’ is
presented with respect to the depth for five boreholes (BH) as in Figure 3.3. Results
of Standard Penetration Tests were given in detail at Appendix A.

Obtained undisturbed soil samples were subjected to classification tests such as sieve
analysis and consistency limit tests. Grain size distribution curves (Appendix B) and
Atterberg limits of soils taken from different depths were obtained. Then, soil
samples were classified according to Unified Soil Classification System (USCS). As
a result, the site was idealized as following: First 6m of soil strata was consisted of
clayey sand (SC) having plastic and liquid limit of 19% and 28%, respectively. It
was underlain by low to highly plastic clay (CL-CH) layer with a thickness of 9m
resting over very stiff highly plastic clay (CH). It was observed that the ground water
table was located at 5m below the ground surface. According to physical soil
properties and average N60’ values, the soil profile was demonstrated as in Figure
3.4.

Figure 3.4. The soil profile of the site

31
3.1.2. Dynamic properties

Multichannel Analysis of Surface Waves (MASW) tests were used to determine


dynamic properties of the site. MASW utilizes the reflection of waves generated at
ground surface for imaging subsurface (Park et al., 1999). A sledge hammer was
used to generate surface waves by impact (Figure 3.5a). The velocity and arrival time
of those waves were measured by 26 geophones (4.5Hz) emplaced with a linear
spacing of 4m (Figure 3.5b). The tests were carried out for both north-south and east-
west directions as illustrated in Figure 3.2.

(a) (b)

Figure 3.5. MASW test in the field (a) generation of impact (b) geophones

The major part of the waves generated at ground surface are Rayleigh waves (Miller
and Pursey, 1955). Therefore, waves whose characteristics were determined through

(a)

32
(b)

Figure 3.6. Compression wave velocity profile of the site (a) north-south direction (b) east-west
direction

MASW tests, were Rayleigh waves, as well. Since, Rayleigh waves are comprised of
interference of P-waves and SV-waves, near surface compression (Figure 3.6) and
shear (Figure 3.7) wave velocity profiles of the site can be estimated by inversion of
Rayleigh waves (Xia et al., 1999).

(a)

33
(b)

(c)

Figure 3.7. Shear wave velocity profile of the site (a) north-south direction (b) east-west
direction (c) average

A series of microtremor tests were conducted in order to estimate predominant


period of the site under low amplitude ambient vibrations naturally existed in the
field (Figure 3.8). Vibrations were measured by a 24 bit Sara System SR04S3 triaxial
digital seismometer. The resolution of the sensor was 2 nm/s. Measurements were

34
performed with sampling rate of 100 Hz. Consequently, the results showed that the
predominant period of the site was approximately 0.32 s.

(a) (b)

Figure 3.8. Microtremor test (a) data acquisition in the field (b) SR04S3 triaxial seismometer

3.2. Test Equipment

3.2.1. Vibration source

As can be understood from section 2.1, most of the vibration isolation problems deals
with continuous vibration sources. In previous studies based on field experiments, it
was seen that vibratory compactors and mechanical oscillators were used as
continuous sources in the field. In this study, plate vibratory compactor was used to
originate continuous vibrations on the surface (Figure 3.9).

Figure 3.9. Plate vibratory compactor

35
It was intended to simulate vibrations caused by traffic and construction activities
which typically generate vibrations in a frequency range of 10-60 Hz (Henwood and
Haramy, 2002) with centrifugal force of 10-200 kN (Day and Benjamin, 1991).
These vibrations result in peak particle velocity of 1-50 mm/s in the ground
(Henwood and Haramy, 2002; Anonymous, 2004). Plate vibratory compactor used in
the current study, has a 4hp gasoline engine and base plate with dimensions of 30cm
(width) x 48cm (length) and weights 58 kg in total (including base plate). It can
generate vibrations in a frequency range of 20-80 Hz. Thus, vibratory compactor was
employed to excite frequencies of around 25 Hz, 50 Hz and 70 Hz (look section 3.3
for further information) by adjusting operating speed. Under these circumstances, it
induced 3mm/s of peak particle velocity in the ground by centrifugal force of 10 kN.
In brief, the plate vibratory compactor was appropriate to simulate vibrations due to
traffic and construction activities.

3.2.2. Accelerometer

Ground-borne vibrations were measured by Sense Box 7001 accelerometers (Figure


3.10) which could operate in uniaxial direction. They are MEMS (micro electro-
mechanical system) type nitrogen damped sensors designated for field tests
conditions. They have frequency bandwidth of 0-400 Hz in acceleration range of -2
to +2 g. Sensitivity of the accelerometers is 5μg which can perform measurements at
very low noise levels. Six accelerometers were located at specific locations
(described at section 3.3) to measure the accelerations in vertical direction.

Figure 3.10. MEMS type Sense Box 7001 accelerometer

36
3.2.3. Data acquisition system

Outputs of accelerometers were digitized by Test Box 2010 dynamic data logger
(Figure 3.11). The system has resolution of a 24 bit ADC (analog to digital
converter). It provides simultaneous sampling within a rate range of 1 Hz to 2 kHz.
This 8-channel dynamic data acquisition system needs input voltage of ±12V. The
energy was supplied by Tuncmatik Newtech Pro UPS (uninterruptible power
supply). The UPS provided 3 kVa with sinusoidal excitation during experiments.

Figure 3.11. Test Box 2010 dynamic data logger

3.3. Field Test Procedure

Vibration isolation performance of trench type wave barriers was intended to be


measured by field tests. First, vibration tests were carried out in the absence of wave
barriers. In that way, attenuation of vibrations were observed by the effect of
radiation and material damping only. Later, a series of vibration tests were conducted
with presence of different wave barriers such as open trench, EPS geofoam wave
barrier and water filled trench. Then, screening effectiveness of the barriers were
determined by comparing the isolation performance of the barriers with those
obtained in the no-trench condition.

The test site had majorly smooth surface, but some slightly rough planes were
encountered as well. Prior to vibration tests, the ground surface was flattened by an
earth mover for leveling the vibration source and accelerometers (Figure 3.12).

37
Figure 3.12. The location of accelerometers in the field after levelling

A plate vibratory compactor operating within the range of 20 Hz to 80 Hz (section


3.2.1), was used in the field to generate continuous vibrations. The excitation
frequencies were selected as 25 Hz, 50 Hz and 70 Hz considering the normalized
depth of the barriers to be 0.5, 1.0 and 1.5 (described at section 3.3.1).

Two different test configurations were designated to distinguish the influence of the
vibration source-wave barrier distance on screening effectiveness. Vibrations were
generated with frequencies of 25 Hz, 50 Hz and 70 Hz for two different location
plan. Since vibrations were measured in terms of acceleration, accelerometers were
deployed as shown in Figure 3.13.

38
Figure 3.13. Plan view of field test configurations (a) when source-barrier distance (lt) was 2.5m
(b) when source-barrier distance (lt) was 5.5m

Acceleration sensors were named from #1 to #6. Induced vibrations were measured
by accelerometers in vertical direction without presence of any wave barrier. The
data obtained from accelerometers was measured in terms of voltage.

Accelerometers had been connected to the dynamic data logger (section 3.2.3). The
data acquisition system simultaneously transmitted the obtained voltage values to a
computer via an ethernet cable. Acceleration measurements were recorded with the
sampling rate of 200 Hz by using a software named Easy Test-Network. The
interface of the program enabled storage of the measurement data by creating a text
file coded in ANSI. Consequently, acceleration amplitudes obtained at all sensors
were saved in different text files for each measurement series.

3.3.1. Open trench

It was aimed to investigate the screening effectiveness of open trenches. According


to previous studies, vibration isolation performance of open trenches mainly depends
on the geometry of the barrier (section 2.3.1). Therefore, dimensions of the open
trench were determined at first. The most important parameter was the depth of the
barrier (section 2.4.4). So as to acquire maximum performance from the open trench,
it was intended to construct the deepest trench by considering the physical properties
of the site. In that the ground water table was located at 5m below the ground
surface, the depth of the barrier was selected as 4.5m. It was seen in the literature the
vibration isolation performance of the trench was independent from its width (section
2.4.2) and length (section 2.4.3). Since the width of the excavator was 80 cm, the

39
width of the trench was determined as 80 cm for practical reasons. In order to
provide plane stress conditions, length was determined as greater than 7 times of the
width at least. The length of the barrier was selected as 6m (0.8x7=5.6m<6.0m).

Previous studies showed that Rayleigh wave length (λR) had influence on vibration
isolation. Therefore, Rayleigh wave characteristics of the site was estimated.
Equation 2.6 showed that Rayleigh wave velocity was averagely 0.9 times of the
shear wave velocity. Average S-wave velocity at ground surface was observed as
236.5 m/s from Figure 3.7c. As a result, Rayleigh wave velocity was calculated as
212.85 m/s. Rayleigh wave length can be calculated by using Eq. 2.7. The wave
length depends on wave velocity of the soil and excitation frequency of the vibration
source. Since the plate vibratory compactor was operated under different frequencies,
different Rayleigh wave lengths were obtained as shown in Table 3.1. In order to
evaluate the effect of

Figure 3.14. Illustration of dimensionless geometrical properties of the barrier

Rayleigh wave length and the dimensions of the barrier at the same time, barrier’s
geometrical properties were normalized with respect to Rayleigh wave length.
Dimensionless geometrical properties were illustrated in Figure 3.14. Normalized
depth, width and vibration source-wave barrier distance were denoted by D, W and
L, respectively. The variation of normalized dimensions with respect to Rayleigh
wave length was shown in Table 3.1.

40
Table 3.1. Normalized dimensions of the barrier with respect to Rayleigh wave length

L=lt/λR
Frequency VR
λR (m) D=dt/λR lt=5.5 W=wt/λR
(Hz) (m/s) lt=2.5 m
m
25 212.85 8.51 0.53 0.29 0.65 0.09
50 212.85 4.26 1.06 0.59 1.29 0.19
70 212.85 3.04 1.48 0.82 1.81 0.26

Subsequent to vibration tests conducted in the absence of wave barriers, open trench
with a depth of 4.5 m, width of 80 cm and length of 6 m was excavated (Figure
3.15). Since, first 6m of the soil strata was clayey sand, open trench didn’t collapse
due to instability. Thus, no siding or casing was employed to construct the open
trench. Then, vibrations were generated with frequencies of 25 Hz, 50 Hz and 70 Hz.
For two different trench locations (lt=2.5 m and lt=5.5 m), acceleration values were
recorded as described in section 3.3.

(a) (b)

Open trench

Figure 3.15. Open trench in the field (a) general layout (b) closer view

3.3.2. In-filled trenches

Owing to the fact that first 4.5m of the soil strata was clayey sand (SC), no sliding or
failure was observed in the excavation of the open trench. However, construction of
the open trench in different soils having only cohesionless content is not possible
without using any siding or fill material. Thus, the structural instability possibility of
open trenches promoted using in-filled wave barriers. Herewith, EPS geofoam and
water were employed as fill materials in open trenches.
41
3.3.2.1. EPS geofoam filled trench

EPS geofoam is lightweight material with a density range of 16-32 kg/m3. Thus, it
was used as a soft wave barrier in vibration isolation systems (section 2.3.2.1). In the
current study, EPS geofoam sheets were employed to fill open trench. The density of
the geofoam was 16 kg/m3. First of all, dynamic characteristics of the EPS geofoam
was determined. In an attempt to find out the compressional wave velocity (VP) of
the EPS, a series of Ultrasonic Pulse Velocity Test was carried out (Figure 3.16a). A
cubic geofoam sample was prepared with a length of 5 cm. EPS sample was located
between Proceq transducers. The travel time of the pulse along the geofoam was
measured and the length of the sample was divided by the time. This test was
repeated for several times. As a result, compressional wave velocity of the EPS
geofoam was averagely derived as 750 m/s. In order to obtain shear wave velocity of
the geofoam, a series of vibration impact tests were conducted in the laboratory
(Figure 3.16b). EPS geofoam sheets were emplaced and two accelerometer were
deployed on them. Then, an impact was generated on EPS sheets. Knowing the
distance between the accelerometers, the arrival time of the S-wave was obtained.
Consequently, shear wave velocity of the EPS geofoam was obtained as 350 m/s.

(a) (b)

Figure 3.16. Determination of dynamic characteristics of EPS geofoam (a) Ultrasonic Pulse
Velocity Test (b) laboratory impact test

Each EPS geofoam sheets had dimensions of 5cm (height) x 50cm (width) x 100cm
(length). After the completion of field vibration tests with open trench, geofoam
sheets were placed in the trench. The plate vibratory compactor and the
accelerometers were located as in Figure 3.13 (for both lt=2.5m and lt=5.5m). Then,
the vibrations were generated with frequencies of 25 Hz, 50 Hz and 70 Hz.
42
Acceleration amplitudes at specific points were measured in vertical direction and
recorded as explained in section 3.3.

Figure 3.15. EPS geofoam filled trench in the field

3.3.2.2. Water filled trench

Another fill material used in the open trench was water. Since first 4.5 m of the soil
profile was clayey sand (SC), permeability of this layer was very low. Thus, the
trench was directly filled with water without any water proof protection such as
membrane (Figure 3.18). After filling the trench with water, the plate vibratory
compactor and the accelerometers were located as in Figure 3.13 (for both l t=2.5m
and lt=5.5m). Then, the vibrations were generated with frequencies of 25 Hz, 50 Hz
and 70 Hz. Acceleration amplitudes at specific points were measured in vertical
direction and recorded as reported in section 3.3.

43
Figure 3.18. Water filled trench in the field

3.4. Calibration and Filtering of Data

3.4.1. Calibration

Accelerometers measured vibrations in terms of voltage. Therefore, the measured


voltage values were transformed into acceleration amplitudes in g (gravitational
acceleration). All six accelerometers were identical. In a sensor, 2.0 V indicated 1.0 g
and -2.0 V indicated -1.0 g, as well. Accordingly, voltage values were multiplied by
a scale factor of 0.5. Then, obtained data were written in text files coded in ANSI.
Since vibration having three different frequencies were analyzed in case of no-
trench, open trench, EPS geofoam wave barrier and water filled trench for two
different trench locations, 24 acceleration-time history files were created. Each file
contained acceleration amplitudes of six accelerometers measured in time domain.

3.4.2. Filtering of Data

Vibrations were generated with frequencies of 25 Hz, 50 Hz and 70 Hz.


Nevertheless, unwanted frequency content may also have been recorded during
44
signal processing procedure. Therefore, Seismo Signal V5.1.0 program was used to
filter obtained data. The software enables user to filter noise in the given signal,
constitute response spectra and Fourier spectrum of the accelerograms and derive
characteristic properties of the input motion.

(a)

(b)

(c)

Figure 3.19. Application of 8-order Butterworth filter (a) f=25 Hz (b) f=50 Hz (c) f=70 Hz

Butterworth Filter (Butterworth, 1930) was preferred to filter the motion within the
frequency range. The Butterworth Filter is a linear analog electronic filter type used
in signal processing. The filter was designed to screen out the undesired frequency
content out of passband. The frequency range of filter was determined as ±5 Hz.
Average gain at each frequency level was demonstrated in Figure 3.19. Thus and so,
raw data was filtered by 8-order linear Butterworth Filter and acceleration-time
history files for each vibration test series were established.

45
4. RESULTS AND DISCUSSION

Vibrations generated with frequencies of 25 Hz, 50 Hz and 70 Hz, were measured by


six accelerometers at specific locations in case of no trench, open trench EPS
geofoam and water filled trench. Acceleration-time histories were obtained at each
accelerometer. In order to analyze the variation of vibrations with distance, peak
particle acceleration amplitudes were derived at each point. An example for
derivation of maximum acceleration amplitude is shown in Figure 4.1. The procedure
was repeated for every field vibration test series.

46
Figure 4.1. Obtaining peak particle acceleration amplitudes (in case of no trench under
vibrations having frequency of 25 Hz) from accelerograms (a) #1 (b) #2 (c) #3 (d) #4 (e) #5 (f) #6

4.1 Attenuation of Vibrations

Vibrations were generated by using a plate vibratory compactor. Even though the
frequencies of the vibrations were same for identical tests, the input motion could
have some slight differences since the vibratory compactor could not generate same
vibrations in every vibration test. Thus, the peak particle acceleration values obtained
at each accelerometer, were normalized with those obtained in the first sensor
(accelerometer #1). The variation of normalized acceleration values with the distance
was illustrated for different trench locations in time domain. (Figure 4.2 and 4.3). In
addition, the influence of the dimensionless trench depth was also regarded in

47
attenuation of ground-borne vibrations for absence (no trench) and presence of wave
barriers such as open trench, EPS geofoam wave barrier and water filled trench.

Figure 4.2. Variation of vibration attenuation when l t= 2.5m (a) no trench-open trench (b) no
trench-EPS geofoam filled trench (c) no trench-water filled trench

48
Figure 4.3. Variation of vibration attenuation when l t= 5.5m (a) no trench-open trench (b) no
trench-EPS geofoam filled trench (c) no trench-water filled trench

In Figure 4.2 and 4.3, the continuous lines indicate free field (no trench) case, the
dashed lines are demonstration of presence of wave barriers. In the absence of wave
barriers, vibrations attenuated in the influence of radiation (material) and material
damping only. The amplitudes of normalized accelerations decreased in an
49
exponential trend by increasing distance. On the whole, approximately 90% of the
dispatched vibrations disappeared at a distance of 10 m. The normalized acceleration
amplitudes decreased by 95% after 15 m away from the first accelerometer.
Accordingly, normalized acceleration values recorded at sensor #5 and #6, were very
close to each other. In the region of 15 m away from the source, accelerations were
measured at a level of 10-4-10-5 g. Since the sensitivity of the accelerometers was
approximately 5μg, the measurements around this region were very sensitive to noise
fluctuation. Consequently, it was more accurate to use the data obtained at
accelerometers #3 and #4 when determining screening effectiveness of wave barrier.

In Figure 4.3, it was observed that the rate of attenuation increased with higher
excitation frequencies in case of no trench. As vibration frequency increased, the
amplitudes of normalized acceleration values attenuated in a more rapid manner.
Results of microtremor tests showed that predominant period of the site was 0.32 s.
In other words, the frequency of the site was about 3 Hz. Since soil could transmit
vibrations having frequencies close to that of the soil without too much dissipation,
transmissibility of the soil decreased with increasing vibration frequency. The reason
of this observation may be that transmissibility of the soil decreased at higher
frequency (than that of soil) levels.

Acceleration amplitudes were expected to decrease with increasing distance from the
vibration source. However, some amplifications in acceleration values were observed
at some points such as point #3 in case of no trench under vibration frequency of 70
Hz when lt=2.5 m (Figure 4.2), point #3 in case of no trench under vibration
frequency of 50 Hz and 70 Hz when lt=5.5 m (Figure 4.3). Haupt (1981), Al Hussaini
(1992) and Massarch (2005) mentioned about similar phenomenon. Haupt (1981)
indicated that distinct peaks could be observed when shear wave velocity of a site
increased in proportional with the depth. In addition, Al Hussaini (1992) noted that
this phenomenon could occur in layered soil profiles due to reflection of the waves at
interfaces of soil layers. Massarsch (2005) stated that surface waves and reflected
waves could superpose in same phase and originate a bigger wave which could cause
amplification in acceleration. Since the soil profile in the current study had layered
profile and increasing shear wave velocity profile, the unexpected distinct peaks may
be explained by superposition of surface waves and reflected waves in same phase.

50
4.2. Vibration Isolation by Wave Barriers

In the presence of wave barriers (open trench, EPS geofoam wave barrier and water
filled trench), normalized acceleration values were located under those of obtained in
the case of no trench (Figure 4.2 and 4.3). It indicated vibration isolation due to wave
barriers. Attenuation of vibrations was analyzed in terms of acceleration in time
domain by now. In order to investigate the screening effectiveness of the wave
barriers, Fourier spectrums of the vibrations were constructed by using Seismo
Signal V.5.1.0 software (Figure 4.4). The data obtained from accelerometers were
transformed into spectral amplitude in frequency domain. Maximum spectral
amplitudes in each Fourier spectrum were normalized with those obtained from first
sensor (Figure 4.5 and 4.6).

51
Figure 4.4. Obtaining Fourier Spectrums (in case of 4.5m deep open trench under vibrations
having frequency of 50 Hz) from accelerograms (a) #1 (b) #2 (c) #3 (d) #4 (e) #5 (f) #6
52
Figure 4.5. Attenuation of spectral amplitudes when l t= 2.5m (a) no trench-open trench (b) no
trench-EPS geofoam filled trench (c) no trench-water filled trench

53
Figure 4.6. Attenuation of spectral amplitudes when l t= 5.5m (a) no trench-open trench (b) no
trench-EPS geofoam filled trench (c) no trench-water filled trench

Woods (1968) proposed to use an efficiency ratio to evaluate vibration isolation


performance of the wave barriers quantitatively. The efficiency ratio was called as

54
amplitude reduction ratio and defined as the ratio between peak spectral amplitudes
of vibrations obtained before and after implementation of the barrier (Eq. 4.1).

(A max ) after
AR  (4.1)
(A max ) before

In the present study, the amplitude reduction ratios at measurement points were
calculated by dividing the normalized Fourier amplitudes obtained in presence of
wave barriers by those obtained in the absence of any wave barrier. Since vibration
isolation was expected behind the trench, this procedure were carried out for
accelerometers located at points #3, #4, #5 and #6. The amplitude reduction ratios
obtained in presence of wave barriers were analyzed with respect to the excitation
frequency and the normalized trench depth. Screening efficiency of wave barriers
were examined in terms of excitation frequency of the vibrations, normalized trench
depth, distance from vibration source and vibration source-wave barrier distance.
Since vibrations were generated in a frequency range of 25 Hz – 70 Hz, Rayleigh
wave length changed between 3.04 m and 8.51 m. Thus, normalized trench depth
was obtained in a range of 0.5-1.5 (Table 3.1). Therefore, the influence of excitation
frequency and normalized trench depth considered at the same time.

4.2.1. Screening efficiency of open trenches

In order to examine vibration isolation performance of open trenches, amplitude


reduction ratios were calculated for two different trench locations with varying
normalized trench depth. The variation of amplitude reduction ratio with distance is
illustrated at Figure 4.7.

Normalized trench depth varied between 0.5 and 1.5 depending on the excitation
frequencies of vibrations. Amplitude reduction ratios were about 1.0 for D=0.5 or
f=25 Hz which meant that no reduction was obtained. For D≥1.0, the reduction ratios
were generally located under the straight line of AR=1. It can be noted that the trench
depth for open trench should be greater than 1.0λR to achieve reasonable isolation.

55
Figure 4.7. Variation of amplitude reduction ratio for open trench when (a) l t= 2.5m (b) lt= 5.5m

When the open trench was 2.5m away from the vibration source and D≥1.0,
amplitude reduction ratios at accelerometer #3 and #4, were in the range of 0.02 to
0.33 (Figure 4.7a). Reduction ratio values obtained from accelerometer #5 and #6
were higher than those of obtained from accelerometer #3 and #4. This situation was
observed when the trench was 5.5 m far from the source (Figure 4.7b), as well. Thus,
it may be stated that amplitude reduction ratios increases with increasing distance
from the wave barrier. More reduction can be achieved in a region close to the open
trench.

Accelerometer #5 in case of lt=2.5 m, was located 11.8 m far from the vibration
source. When the open trench was constructed 5.5 m away from the source,
accelerometer #4 was deployed at a distance of 10.8 m. Accordingly, they were
placed at approximately same vicinity. In order to study the influence of vibration
56
source-open trench distance on screening effectiveness, the amplitude reduction
ratios obtained at these accelerometers were compared. Since no reduction was
observed when normalized trench depth was 0.5, comparison was performed for
D≥1.0. The amplitude reduction ratio was obtained averagely 0.31 for lt=2.5 m. In
case of lt=5.5 m, the ratio was calculated in a range of 0.28 and 0.33. Same
comparison was performed between accelerometer #4 in case of lt=2.5 m and
accelerometer #3 in case of lt=5.5m since both of the sensors were located in a region
6.8-7.8 m far from the barrier. Amplitude reduction ratios were obtained
approximately 0.1 and 0.15, respectively. It may be stated that approximately same
amplitude reduction values were obtained in same vicinity independent from the
location of wave barrier. However, all dimensionless spectral amplitudes were
normalized with those of obtained in the first accelerometer prior to calculating
amplitude reduction ratios. Amplitude of vibration depends on combination of
surface wave and reflected wave in-phase or out-phase conditions. As the distance
between vibration source-wave barrier decreases, reflected waves become more
determinant. Therefore, acceleration value in front of the barrier may have been
measured more or less than it was supposed to be. As a result, measurements in this
region were susceptible to this kind of disparity. Due to measurement discrepancy in
front of the wave barrier, it is impossible to draw a general conclusion about
influence of vibration source-wave barrier distance on screening effectiveness.

The average of amplitude reduction ratios was calculated with regards to normalized
trench depth for different trench locations. When the open trench was installed 2.5 m
away from the vibration source, average reduction ratios for D=0.5, D=1.0 and
D=1.5 were found as 0.65, 0.19 and 0.18, respectively. In case of lt=5.5 m, the
average values for D=0.5, D=1.0 and D=1.5 were calculated as 1.04, 0.32 and 0.64,
respectively. The reason of obtaining only 36% reduction when D=1.5 and lt=5.5m
may be noise fluctuations at that accelerometers. Since measured acceleration
amplitudes at point #5 and #6 were at a level of 10-5 g, the sensors were very
sensitive to noise for that level of acceleration. In addition, vibrations were reduced
by more than 95% at that region in the absence of any wave barrier. Thus, reduction
due to open trench was slightly observed. Consequently, it may be summarized that
the open trench with a depth of at least 1.0λR can provide satisfactory level of
vibration isolation.
57
4.2.2. Screening efficiency of in-filled trenches

4.2.2.1. Vibration isolation performance of EPS geofoam wave barriers

The amplitude reduction ratios at measurement points in case of EPS geofoam wave
barrier were calculated as in section 4.2.1. The variation of amplitude reduction ratio
with respect to the distance is illustrated at Figure 4.8.

Figure 4.8. Variation of amplitude reduction ratio for EPS geofoam wave barrier trench when
(a) lt= 2.5m (b) lt= 5.5m

Normalized depth of the EPS geofoam filled trench increased with increase of
vibration frequency. It was observed that amplitude reduction ratio values obtained
in case of D=0.5, were greater than those obtained when D≥1.0. This indicated that
more reduction was achieved at higher normalized barrier depth. In order to examine
58
the influence of normalized depth on screening efficiency of geofoam barrier,
average reduction ratios were found for each vibration frequency level. When the
barrier was constructed 2.5 m away from the vibration source, average reduction
ratios for D=0.5, D=1.0 and D=1.5 were obtained as 0.48, 0.11 and 0.20,
respectively. When lt=5.5 m, the averaged values for D=0.5, D=1.0 and D=1.5 were
calculated as 0.52, 0.21 and 0.24, respectively. Considering average amplitude
reduction values, it may be summarized that all EPS geofoam barriers procure
reasonable efficiency. According to Woods (1968), the amplitude reduction ratio
should have been less than 0.25 for a barrier to be deemed as efficient. However, the
barrier with a depth of 0.5 λR could only reduce 52% of the ground-borne vibrations.
Thus, it is suggested to use EPS geofoam wave barrier having normalized depth of
1.0 λR for sufficient efficiency.

In order to analyze the influence of distance from source on attenuation of vibrations


in presence of geofoam barrier, amplitude reduction ratios were examined in terms of
distance in identical cases. Since efficiency of the barrier having normalized depth of
0.5λR could not provide adequate isolation performance, it was excluded from
comparison. In case of EPS geofoam filled trench located 2.5m away from the
vibration source, the reduction ratios were averagely obtained as 0.025, 0.105, 0.235
and 0.26 at points #3, #4, #5 and #6, respectively. When the barrier was constructed
at a distance of 5.5 m away from the source, the averaged values were obtained as
0.085, 0.2, 0.255 and 0.365. These results show that more reduction can be achieved
in a region close to the barrier as in the open trench case (reasons were explained in
section 4.2.1).

Accelerometer #5 when lt=2.5 m and accelerometer #4 when lt=5.5 m were deployed


at approximately same vicinity (Figure 3.13). In order to examine the effect of
source-barrier distance, the reduction ratios at these points were compared. When
D≥1.0 and lt=2.5 m, average amplitude reduction ratio was 0.235, however it was
obtained as 0.2 when D≥1.0 and lt=5.5 m. Same comparison was performed between
accelerometer #4 in case of lt=2.5 m and accelerometer #3 in case of lt=5.5m since
both of the sensors were located in a region 6.8-7.8 m far from the barrier. Amplitude
reduction ratios were obtained approximately 0.085 and 0.095, respectively. It may
be stated that approximately same amplitude reduction values were obtained in same

59
vicinity independent from the location of wave barrier. Due to aforementioned
(section 4.2.1) discrepancies about measurement of acceleration amplitudes in front
of the barrier, it is still impossible to draw a conclusion about influence of vibration
source-wave barrier distance on screening effectiveness.

4.2.2.2. Vibration isolation performance of water-filled trench

Vibration isolation performance of water filled trench was examined in terms of


amplitude reduction ratio. The variation of the reduction ratios with respect to
distance was shown in Figure 4.9.

Figure 4.9. Variation of amplitude reduction ratio for water filled trench when (a) l t= 2.5m (b)
lt= 5.5m

60
In Figure 4.9a, the variation of reduction ratios was illustrated when the water filled
trench was located 2.5 m far from the vibration source. The points located above the
line of AR=1.0 indicated amplification in vibration in the presence of water filled
trench. No regular pattern was observed in this case. In-filled trenches having
normalized depth of 0.5 and 1.5, had no isolation effect. It may be stated that the
amplification can be caused by superposition of surface waves and reflected waves in
same phase. Vibration isolation was only observed when the trench depth was 1.0λR.
Averagely 76% of vibrations were reduced due to presence of the water filled trench.

When the water filled barrier was constructed 5.5 m far from the vibration source,
amplification in reduction ratios was infrequently observed (Figure 4.9b). The trench
having depth of 0.5λR failed to reduce vibrations. Water filled trench having
normalized depth of 1.0 and 1.5, provided average amplitude reduction ratio of 0.3
and 0.27, respectively. Through no regular pattern about screening effectiveness of
water filled trench was observed, influence of source-barrier distance could not be
investigated.

Vibration isolation performance of water filled trench was not observed in a regular
trend. Though water prevent transition of shear waves, passage of compressional
waves could cause underperformance of the barrier. Since amplification of vibrations
in presence of the water filled trench was often encountered, the location of distinct
peaks needs further investigation for the use of this type barriers.

4.3. Comparison on Screening Effectiveness of Wave Barriers

Screening effectiveness of the open trench, EPS geofoam wave barrier and water
filled trench was analyzed in terms of amplitude reduction ratio. The field tests were
conducted for two different trench locations under vibrations having frequencies
varied between 25 Hz and 70 Hz. Obtained amplitude reduction ratios in each test
series were summarized in Table 4.1 and 4.2.

61
Table 4.1. Obtained amplitude reduction ratios in presence of wave barrier when l t=2.5 m

AMPLITUDE REDUCTION RATIO (AR)


OPEN TRENCH
Point #3 Point #4 Point #5 Point #6 AVERAGE
D≈0.5 (25Hz) 0.28 0.79 0.61 0.91 0.65
D≈1.0 (50Hz) 0.18 0.06 0.30 0.21 0.19
D≈1.5 (70Hz) 0.02 0.13 0.32 0.24 0.18
EPS GEOFOAM FILLED TRENCH
Point #3 Point #4 Point #5 Point #6 AVERAGE
D≈0.5 (25Hz) 0.31 0.28 0.34 0.98 0.48
D≈1.0 (50Hz) 0.02 0.15 0.15 0.13 0.11
D≈1.5 (70Hz) 0.03 0.06 0.32 0.39 0.20
WATER FILLED TRENCH
Point #3 Point #4 Point #5 Point #6 AVERAGE
D≈0.5 (25Hz) 11.04 1.01 1.27 4.91 4.56
D≈1.0 (50Hz) 0.14 0.22 0.29 0.33 0.24
D≈1.5 (70Hz) 0.49 0.63 2.71 2.79 1.65

Table 4.2. Obtained amplitude reduction ratios in presence of wave barrier when l t=5.5 m

AMPLITUDE REDUCTION RATIO (AR)


OPEN TRENCH
Point #3 Point #4 Point #5 Point #6 AVERAGE
D≈0.5 (25Hz) 1.78 0.58 0.89 0.92 1.04
D≈1.0 (50Hz) 0.21 0.33 0.50 0.26 0.32
D≈1.5 (70Hz) 0.08 0.28 1.39 0.82 0.64
EPS GEOFOAM FILLED TRENCH
Point #3 Point #4 Point #5 Point #6 AVERAGE
D≈0.5 (25Hz) 0.58 0.71 0.43 0.37 0.52
D≈1.0 (50Hz) 0.13 0.26 0.21 0.25 0.21
D≈1.5 (70Hz) 0.04 0.14 0.30 0.48 0.24
WATER FILLED TRENCH
Point #3 Point #4 Point #5 Point #6 AVERAGE
D≈0.5 (25Hz) 1.09 3.33 0.85 1.01 1.57
D≈1.0 (50Hz) 0.31 0.39 0.33 0.16 0.30
D≈1.5 (70Hz) 0.14 0.20 0.49 0.25 0.27

According to average amplitude reduction ratio values (Table 4.1 and 4.2), open
trench and geofoam wave barriers having depth of at least 1.0 λR provided sufficient
vibration isolation. Generally, water filled trenches having different normalized
depths failed to reduce ground borne vibrations. Even though water filled trench
having D=0.5 performed well when lt=2.5 m, barrier having same dimensions failed
to attenuate vibration when it was constructed 5.5 m away from the source. Water

62
filled trench performed worst among other wave barriers. Shear waves can not
propagate through water. Since Rayleigh wave was comprised of compressional and
shear wave, the reason of water filled trench to fail may be attributed to the
contribution of compressional waves.

Efficiency of open trench and EPS geofoam filled trench was very close to each
other except open trench having depth of 1.5 λR when lt=5.5 m. In this test, distinct
peak was observed at point #5 and high amplitude reduction ratio was obtained at
point #6. As aforementioned, efficiency of the barriers decreased at long distances
from the vibration source due to low transmissibility of soil, noise fluctuation and
superposition of surface wave and reflected waves. Thus, averaged reduction ratio
for this case was obtained high by the virtue of measurements at point #5 and #6.

Accelerometers #3 and #4 were deployed within in a distance of 4.5 m away the


trench. These points were located near to trench. Thus, average reduction around the
trench was calculated by averaging the AR values in this region (points #3 and #4).
When the open trench was located 2.5 m away from the vibration source, amplitudes
of vibrations were reduced by 88% and 92.5% around the trench for D=1.0 and
D=1.5, respectively. When the open trench filled with EPS geofoam, average values
were found as 91.5% and 95.5% around the trench for D=1.0 and D=1.5,
respectively. When open trench was located 5.5 m away from the source, average
amplitude reduction ratios obtained near to trench, were obtained as 0.27 and 0.18
for D=1.0 and D=1.5, respectively. The average reduction ratios were found as 0.195
and 0.09 in case of geofoam filled trench when D=1.0 and D=1.5, respectively. Two
major conclusions may be derived by these observations. The first one is that more
reduction can be achieved in a region close to the barrier as mentioned in section 4.2.
The influence of wave barrier on vibration isolation decreases with distance from the
wave barrier. The second inference is that the efficiency of the wave barrier increases
with increasing normalized trench depth. Better isolation performance can be
achieved by deeper trenches. The efficiencies of open trench and EPS geofoam filled
trench having same depths are very close to each other.

63
4.4. Comparison between Results of Current Study and Published Studies

The results showed that isolation performance of the barrier mostly influenced by
normalized barrier depth. The effect of vibration source-wave barrier distance was
not noteworthy with regard to vibration isolation (Alzawi and El Naggar, 2009; Qiu
et al., 2014; Saikia, 2014; Ekanayake et al., 2014). Therefore, screening efficiency of
the wave barriers were examined in terms of amplitude reduction ratio which mainly
depends on normalized trench depth.

4.4.1. Experimental studies

Results obtained from field tests (for open and EPS geofoam filled trench) were
compared with those of published in the literature. Since no experimental study about
screening efficiency of water filled trench was encountered, comparison could not be
performed.

4.4.1.1. Open trench

Woods (1968), Haupt (1981) and Alzawi and El Naggar (2011a) investigated the
isolation performance of open trenches experimentally. Woods (1968) conducted
field experiments with open trenches located on uniform silty fine sand (SM). The
distance of the trench to the vibration source was in the range of 0.2λR-0.9λR. Haupt
(1981) carried out a series of laboratory experiments on open trenches located in 10m
x10m rectangular bin filled with dense sand. The trench was positioned in a distance
range of 0.33 λR and 0.875 λR away from the source. The amplitude reduction ratios
were calculated within a region close (1.5 λR) to the wave barrier. Alzawi and El
Naggar (2011a) performed full scale field test series using open trench located 5 m
away from the vibration source. Measurements were carried out within the range of
5.5-18m. The average amplitude reduction ratios were obtained at different
measurement points.

In the present study, amplitude reduction ratios obtained at points #3 and #4 were
employed in comparison since Haupt (1981) reported the screening effectiveness of
open trench by values obtained near to trench. In the present study, the open trench

64
was located at a distance of 0.29 λR and 1.81 λR from the vibration source. The
comparison between results of present study and those of obtained from experimental
investigations were demonstrated in Figure 4.10.

Figure 4.10. Comparison of amplitude reduction ratios obtained from present study with those
of reported in experimental studies in the literature for case of open trench

In accordance with Figure 4.10, results obtained in the present study seemed to be
generally in a good agreement with those of reported in the literature. However, the
amplitude reduction ratio obtained in the presence of open trench having depth of 0.5
λR fell outside the range of data given in the literature. This observation may be
attributed to superposition of surface waves and reflected wave in same phase for
that vibration frequency. In general, the amplitude reduction ratios were obtained
within the range of 0.1-0.4 for the case of open trench based on previous studies and
present study.

4.4.1.2. EPS geofoam wave barrier

Alzawi (2011) conducted full scale field experiments to examine the screening
effectiveness of EPS geofoam wave barriers. Normalized depth (D) was changed
between 0.57 and 0.84 and normalized distance to the vibration source was in the
range of 0.92-1.36. Variation of the amplitude reduction ratios with respect to the
distance was obtained for different barrier depths. The reported reduction ratios were
compared with those of obtained in the present study.
65
In the present study, geofoam filled trench was located two different places (lt=2.5 m
and lt=5.5 m). Normalized depth of the geofoam filled trench was in the range of
D=0.5-1.5. Influence of the vibration source-wave barrier distance on screening
effectiveness wasn’t observed within source-barrier distance range of 2.5-5.5m.
However, results obtained at different trench locations were compared separately
with those of reported in the literature for the purpose of lucidity.

Figure 4.11. Comparison of amplitude reduction ratios obtained from present study with those
of reported in experimental studies in the literature for case of EPS geofoam filled trench (a)
lt=2.5 m (b) lt=5.5 m

66
In Figure 4.11, the variation of amplitude reduction ratios with respect to the distance
was illustrated for different normalized barrier depths. Results obtained from present
study and Alzawi (2011) were in a good agreement regardless of the distance
between vibration source and wave barrier. It may be deduced from the results that
more reduction can be achieved in a region close to the barrier. The effect of
geofoam filled trench decreased by distance.

Apart from some discrepancies in the reduction ratios obtained by Alzawi (2011),
averagely 50% of vibration amplitudes decreased in the presence of geofoam
barriers. It may be observed that more reduction achieved in a region close to the
barrier. The effect of the geofoam filled trench decreased by distance. Apart from
some discrepancies in the reduction ratios obtained by Alzawi (2011), averagely 50%
of vibration amplitudes decreased in the presence of geofoam barriers. Better
isolation was achieved at the present study in case of EPS geofoam filled trench
having depth of at least 1.0 λR.

4.4.2. Numerical investigations

Results obtained from the present study were compared with those of reported in
previous numerical investigations regarding the use of open trench as wave barrier.
Dolling (1965), Beskos et al. (1986), Al Hussaini (1992), Tsai and Chang (2009),
and Saikia and Das (2014) used numerical methods to study screening effectiveness
of open trench. Beskos et al. (1986), Al Hussaini (1992) and Tsai and Chang (2009)
examined isolation performance of open trenches by using boundary element
method.

Beskos et al. (1986) reported amplitude reduction ratios for different trench
dimensions. The results obtained for normalized trench width of 0.1 and 0.4 were
utilized in comparison since normalized width of the open trench in the present study
varied between 0.09-0.26. Al Hussaini (1992) employed open trench located 5λR
away from the vibration source. The width range was between 0.1λR and 0.3λR in
that study. Tsai and Chang (2009) reported amplitude reduction ratios for open
trench when L=2. Saikia and Das (2014) used finite element method to investigate
the screening efficiency of open trench located 1.0 λR far from the vibration source.

67
As aforementioned in section 4.4.1.1, amplitude reduction ratios obtained in near to
trench region were used in comparison. The comparison between results of present
study and those of obtained by numerical studies were demonstrated in Figure 4.12.

Figure 4.12. Comparison of amplitude reduction ratios obtained from present study with those
of reported in numerical studies in the literature for case of open trench

The results obtained in the present study were generally in a good agreement with
those of reported in the literature. However, open trench having depth of 0.5 λR
underperformed when compared with those in the literature. Reduction ratios
reported by Tsai and Chang (2009) were less than those predicted in other numerical
studies. It may be considered as upper limit of amplitude reduction ratios. Results of
Beskos et al. (1986) and Al Hussaini (1992) may be regarded as lower boundary for
estimation of amplitude reduction. Average amplitude reduction ratios obtained in
the current study was mostly converged by Saikia and Das (2014). It may be used for
estimation of vibration isolation for case of open trench in preliminary design stage.

68
5. SUMMARY AND CONCLUSIONS

5.1. Summary

The aim of the present study is to evaluate the screening effectiveness of trench type
wave barriers by field tests. A great deal of researches have been carried out by
employing numerical methods. Results of the numerical investigations should be
endorsed by field tests. However, number of studies based on field experiments are
very limited. Therefore, the present dissertation attempted to compensate this gap by
field tests.

First of all, a detailed site investigation was carried out to determine physical and
dynamic properties of the site. Governing parameters for screening effectiveness of
wave barriers were determined as excitation frequency of vibration source, trench
dimensions, vibration source-wave barrier distance and fill material. In the field,
continuous vibrations were generated by a plate vibratory compactor in frequencies
of 25 Hz, 50 Hz and 70 Hz. In order to observe attenuation characteristics of
vibration under solely soil damping, the amplitude of dispatched vibrations were
measured by accelerometers at specific points in the absence any barrier. Then,
measurements were performed for case of open trench, geofoam wave barrier and
water filled trench. In order to examine screening effectiveness of wave barriers
quantitatively, amplitude reduction ratio was employed in frequency domain.
Vibration isolation performance of trenches was evaluated regarding vibration
frequency, normalized barrier depth, trench location and fill material. Obtained
results were compared with those of reported in the literature.

5.2. Conclusions

A full scale field vibration test series was conducted to investigate vibration isolation
performance of open and in-filled trench type wave barriers. Major conclusions

69
based upon the evaluation of data obtained from field experiments are summarized as
given below:

1) Attenuation of ground-borne vibrations were examined in the absence and


presence of wave barriers. The results showed that amplitudes of vibrations
decreased more rapidly as excitation frequency of the vibration source
increased. Since predominant frequency of the site was approximately 3 Hz,
the soil could transmit vibrations having frequency level close to 3 Hz
without too much decrease in amplitude. As the excitation frequency
increased, transmissibility ratio of the soil increased. Therefore, the reason of
more rapid decrease in amplitude of vibrations may be low transmissibility
ratio and high damping ratio of the soil at higher frequency (than soil) levels.

2) It is expected that amplitude of vibrations decreases by distance from the


source. However, it was observed that some distinct peaks occurred in the
attenuation curves. Since elastic waves may diffract and reflect at layer
boundaries in layered soil medium, amplitude of waves governed by
superposition of surface waves and reflected and diffracted waves. Therefore,
this phenomena may be attributed to in-phase or out-phase combinations of
superposed elastic waves.

3) The effect of vibration frequency is considered in terms of Rayleigh wave


length. Thus, normalized parameters (with respect to Rayleigh wave length)
were employed in evaluation of isolation performance of wave barriers. It is
seen that Rayleigh wave length and normalized depth of wave barrier have a
major influence on the vibration isolation performance of trench type wave
barriers.

4) Open trench and EPS geofoam wave barrier outperformed water filled trench
in reduction of ground-borne vibrations.

a) Average amplitude reduction ratios in case of open trench having


depth of 0.5 λR was obtained as 0.65 and 1.04. When trench depth was
greater than 1.0 λR, average values were reported in the range of 0.18-
0.64. Reason of higher amplitude reduction ratio may be occurrence
of distinct peaks.

70
b) EPS geofoam filled trench provided 50% reduction in amplitude of
vibrations when D=0.5. Approximately 80% of vibrations was
isolated in presence of geofoam wave barrier having depth of at least
1.0 λR.

c) Vibration amplified in case of water filled trench having normalized


depth of 0.5. At deeper barriers, obtained average amplitude reduction
ratios infrequently varied regardless of trench depth. Thus, isolation
performance of water filled trench was not observed in a regular
trend.

d) In brief, screening efficiency of open trench and EPS geofoam filled


trench is close to each other. Open trench creates discontinuity in soil
medium for elastic waves not to pass. EPS geofoam has very low
density compared with those of soil, thus a part of vibrations may be
isolated while propagating through geofoam media. These may be
reasons of higher efficiency of open trench and geofoam filled trench.
Even though water prevents shear waves to propagate, passage of
compressional waves reduce the screening effectiveness of water
filled trench.

5) Influence of vibration source-wave barrier distance on isolation performance


of wave barrier was examined. Amplitude reduction values obtained in same
vicinity were compared for different trench locations. The results showed that
approximately same amplitude reductions ratios were obtained in same
vicinity independent from the location of the trench. However, as the distance
between vibration source and wave barrier decrease, reflected waves become
more determinant. Amplitude of vibrations governed by superposition of
surface wave and reflected waves. Since dimensionless amplitudes of
vibrations were normalized with those of obtained in front of the trench,
measurements may contain some disparity. As a result, it is impossible to
draw a general conclusion about influence of vibration source-wave barrier
distance on screening effectiveness of wave barriers.

6) Screening efficiency decreased with increasing distance from trench. In a


region close to trench, higher isolation performance was achieved regardless
71
of wave barrier type. This may be attributed to noise fluctuations or
superposition of surface wave and reflected waves at distances away from the
barrier.

7) Amplitude reduction ratios obtained for the cases of open trench and EPS
geofoam filled trench were in a good agreement with those of reported in the
literature, respectively.

a) Among experimental and numerical studies on open trenches, Tsai


and Chang (2009) and Beskos et al. (1986) may be used as upper and
lower boundaries in prediction of amplitude reduction ratios,
respectively. Findings of Saikia and Das (2014) were the closest
prediction to average of field test results. It may be used for
estimation of amplitude reduction in preliminary design stage.

b) Variation of amplitude reduction with respect to distance in case of


geofoam filled trench was compared with that of reported by Alzawi
(2011). The results were consistent with each other.

As a result, open trench and EPS geofoam wave barrier can be used for vibration
isolation purpose. The influence of Rayleigh wave length and trench depth should be
taken into consideration. Use of wave barrier having depth of at least 1.0 λR is
recommended for sufficient screening. Better isolation performance may be achieved
by deeper wave barriers. Due to the instability of open trench, EPS geofoam can be
used as fill material. Even though higher isolation may be provided around a region
close to trench, location of distinct peaks should be considered regarding area to be
protected from vibrations. In conclusion, results obtained in the present study were
valid for the site where all experiments were conducted.

5.3. Recommendations for Future Work

1) Occurrence of distinct peaks directly influences vibration isolation.


Estimation of location and amplification of distinct peaks needs further
investigation.

72
2) Soil profile may vary from site to site. The aforementioned site profile was
consisted of clayey sand over stiff clay. In order to assess the influence of
layering properties on vibration isolation, further field tests should be carried
out in different sites.

3) In the present study, vibration sources such as construction activities and


traffic were employed in the frequency range of 25 Hz-70 Hz. The screening
effectiveness of the wave barriers may be examined for different vibration
sources having other frequency levels.

4) The influence of trench location on screening effectiveness should be


investigated solely. Due to difficulties in measurement of vibration
amplitudes in front of the wave barrier, special attention should be paid to
this issue.

5) In the present study, no casing or siding was employed in the construction of


open trench. The effect of different siding materials such as bentonite or
concrete may be examined by field tests.

6) Open trench, EPS geofoam filled trench and water filled trench was employed
as wave barriers in this dissertation. No experimental study about isolation
performance of stiff barriers was encountered in the literature. Thus,
screening effectiveness of stiff barriers such as sheet pile needs to be
investigated by field test.

73
REFERENCES

Adam, M, and von Estorff, O. (2005) Reduction of train-induced building vibrations


by using open and filled trenches, Comput Struct, 83: 11-24.

Adam, M.A. (2002) Reduction of train induced building vibrations using open and
in-filled trench barriers, 4th International Conference on Civil and
Architecture Engineering, 14-16 May, Cairo, 19 pages.

Ahmad, S. and Al-Hussaini, T.M. (1991) Simplified design for vibration screening
by open and in-filled trenches, J Geotech Eng-ASCE, 117: 67-88.

Ahmad, S., Al Hussaini, T.M. and Fishman, K.L. (1996) Investigation on active
isolation of machine foundations by open trenches, J Geotech Eng-ASCE,
122(6): 454-461.

Al Hussaini, T.M. (1992) Vibration isolation by wave barriers, Ph.D. Thesis, State
University of New York at Buffalo, New York, 212 pages.

Al Hussaini, T.M. and Ahmad, S. (1996) Active isolation of machine foundations by


in-filled trench barriers, J Geotech Eng-ASCE, 122: 288-294.

Alzawi, A. (2011) Vibration isolation using in-filled geofoam trench barriers, Ph.D.
Thesis, The University of Western Ontario, Ontario, 220 pages.

Alzawi, A. and El Naggar, M.H. (2009) Vibration scattering using geofoam material
as vibration wave barriers, 62nd Canadian Geotechnical Conference, 2009,
Halifax, 997-1004.

Alzawi, A. and El Naggar, M.H. (2011a) Full scale experimental study on vibration
scattering using open and in-filled (GeoFoam) wave barriers, Soil Dyn Earthq
Eng, 31: 306-317.

Alzawi, A. and El Naggar, M.H. (2011b) Numerical investigations on vibration


screening, Pan-Am CGS Geotechnical Conference, 2-6 October, Ontario, 8
pages.

74
Amini, Z. (2013) Dynamic characteristics and seismic stability of eps geofoam
embankments, Ph.D. Thesis, The University of Utah, Salt Lake City, 269
pages.

Andersen, L. and Nielsen, S.R.K. (2005) Reduction of ground vibration by means of


barriers or soil improvement along a railway track, Soil Dyn Earthq Eng, 25:
701-716.

Anonymous, Transportation and construction induced vibration guidance manual,


Jones & Stokes, Sacramento, June 2014, 73 pages.

Athanasopoulos, G.A., Pelekis, P.C. and Xenaki, V.C. (1999) Dynamic properties of
EPS geofoam: an experimental investigation, Geosynth Int, 6(3): 171-194.

Aviles, J. and Sanchez-Sesma, F.J. (1983) Piles as barriers for elastic waves, J
Geotech Eng-ASCE, 109(9): 1133-1146.

Aviles, J. and Sanchez-Sesma, F.J. (1988) Foundation isolation from vibrations using
piles as barriers, J Eng Mech-ASCE, 114(11): 1854-1870.

Babu, G.L.S., Srivastava, A., Rao, K.S.N. and Venkatesha, S. (2011) Analysis and
design of vibration isolation system using open trenches, Int J Geomech,
11(5): 364-369.

Barkan, D.D. (1962) Dynamics of bases and foundations, McGraw-Hill Book


Company Inc, New York, 434 pages.

Beskos, D.E., Dasgupta, B. and Vardoulakis, I.G. (1986) Vibration isolation using
open or filled trenches Part 1: 2-D homogeneous soil, Comput Mech, 1:43-63.

Brandenberg, S.J., Nigbor, R.L., Reinert, T., Levulett, D., Stewart, J.P. and Moss,
R.E.S. (2010) Geophysical testing to determine safe vibration limits and
spatial attenuation of vibrations on Sherman Island, NEES-UCLA,
www.nees.ucla.edu/neesrii/ShermanGeophysicalReport, 2010, 14 pages.

Buonsanti, M., Cirianni, F., Leonardi, G., Santini, A. and Scopelliti, F. (2009)
Mitigation of railway traffic induced vibrations: the influence of barriers in
elastic half-space, Advances in Acoustics and Vibrations, 2009:1-7.

75
Butterworth, S. (1930) On the theory of the filter amplifiers, Experimental Wireless
And The Wireless Engineer, 7: 536-541.

Çelebi, E. and Kırtel, O (2013) Non-linear 2-D FE modeling for prediction of


screening performance of thin-walled trench barriers in mitigation of train-
induced ground vibrations, Constr Build Mater, 42: 122-131.

Çelebi, E., Fırat, S., Beyhan, G., Çankaya, İ., Vural, İ. and Kırtel, O. (2009) Field
experiments on wave propagation and vibration isolation by using wave
barriers, Soil Dyn Earthq Eng, 29: 824-833.

Dasgupta, B., Beskos, D.E., and Vardoulakis, I.G. (1990) Vibration isolation using
open or filled trenches Part 1: 3-D homogeneous soil, Comput Mech, 6:129-
142.

Day, D.A. and Benjamin, N.B.H. (1991) Construction Equipment Guide, 2nd Edition,
John Wiley & Sons, Unites States, 464 pages.

Di Mino, G., Giunta, M. and Di Liberto, C.M. (2009) Assessing the open trenches in
screening railway ground-borne vibrations by means of artificial neural
network, Advances in Acoustics and Vibrations, 2009:1-12.

Dolling, H.J. (1968) Schwingungsisolierung von bauwerken durch tiefe, auf


geeignete weise stabilisierte schlitz, Sonderdruck aus VDI Berichte 88,
S.3741.

Ekanayake, S.D., Liyanapathirana, D.S. and Leo, C.J. (2014) Attenuation of ground
vibrations using in-filled wave barriers, Soil Dyn Earthq Eng, 67: 290-300.

El Naggar, M.H. and Chehab, A.G. (2005) Vibration barriers for shock-producing
equipment, Can Geotech J, 42: 297-306.

Erickson, B.A. (2011) Experimental study on the dynamic stress-strain behavior of


expanded polystyrene geofoam using cyclic triaxial tests, MSc Thesis, The
University of Utah, Salt Lake City, 38 pages.

Gao, G.Y., Li, Z.Y., Qiu, C. and Yue, Z.Q. (2006) Three-dimensional analysis of
rows of piles as passive barriers for ground vibration isolation, Soil Dyn
Earthq Eng, 26: 1015-1027.

76
Hamissi, A. and Vileh, R.S. (2007) Optimum geometrical properties of active
isolation by open trenches to reduce vibratory effects of shallow foundations,
International Symposium on Geotechnical Safety & Risk, 18-19 October,
Shangai, 775-784.

Haupt, W.A. (1977) Isolation of vibrations by concrete core walls, 9th International
Conference on Soil Mechanics and Foundation Engineering, 10-15 July,
Tokyo, 2: 251-256.

Haupt, W.A. (1978a) Numerical methods for the computation of steady-state


harmonic wave fields, Dynamical methods in soil and rock mechanics, 1978,
Rotterdam, Balkema, 1: 255-280.

Haupt, W.A. (1978b) Surface waves in non-homogeneous half-space, Dynamical


methods in soil and rock mechanics, 1978, Rotterdam, Balkema, 1: 335-367.

Haupt, W.A. (1981) Model tests on screening of surface waves, 10th International
Conference on Soil Mechanics and Foundation Engineering, 15-19 June,
Stockholm, 3: 215-222.

Haupt, W.A. (1995) Wave propagation in the ground and isolation measures, 3rd
International Conference on Recent Advances in Geotechnical Earthquake
Engineering and Soil Dynamics, 2-7 April, Missouri, 2: 985-1017.

Henwood, J.T. and Haramy, K.Y. (2002) Vibrations induced by construction traffic:
a historic case study, The 2nd Annual Conference on the Application of
Geophysical and NDT Methodologies to Transportation Facilities and
Infrastructure, 15-19 April, Los Angeles, 9 pages.

Horvath, J.S. (1992) New developments in geosynthetics; ‘lite’ products come of


age, ASTM, 20(9): 50-53.

Horvath, J.S. (1996) Geofoam geosynthetic: past, present, and future, EJGE, Invited
paper, 1:1.

Jain, A. and Soni, D.K. (2007) Foundation vibration isolation methods, 3rd WSEAS
International Conference on Applied and Theoretical Mechanics, 14-16
December, Spain, 163-167.

77
Jesmani, M., Hamissi, A., Kamalzare, M. and Vileh, R.S. (2011) Optimum
geometrical properties of active isolation, P I Civil Eng Geotec, 164(6): 385-
400.

Jesmani, M., Shafie, M.R. and Vileh, R.S. (2008) Finite element analysis of active
isolation of deep foundation in clayey soil by rectangular trenches, EJGE,
13(E): 1-16.

Ju, S.H. and Li, H.C. (2011) 3D analyses of open trench barriers filled with water, J.
Geotech. Geoenviron. Eng. 137: 1114-1120.

Kafash, M.H. (2013) Seismic stability analysis of landslides stabilized with expanded
polystyrene (EPS) block geofoam, Ph.D. Thesis, The University of Memphis,
Tennessee, 260 pages.

Kattis, S.E., Polyzos, D. and Beskos, D.E. (1999a) Modelling of pile wave barriers
by effective trenches and their screening effectiveness, Soil Dyn Earthq Eng,
18: 1-10.

Kattis, S.E., Polyzos, D. and Beskos, D.E. (1999b) Vibration isolation by a row of
piles using a 3-d frequency domain bem, Int J Numer Meth Eng, 46: 713-728.

Lamb, H. (1904) On the propagation of tremors over the surface of an elastic solid,
Philos T Roy Soc A, 203: 1-42.

Lu, J.F., Jeng, D.S., Wan, J.W. and Zhang, J.S. (2013) A new model for the vibration
isolation via pile rows consisting of infinite number of piles, Int J Numer
Anal Met, 37: 2394-2426.

Lu, J.F., Xu, B. and Wang, J.H. (2009) A numerical model for the isolation of
moving load induced vibrations by pile rows embedded in layered porous
media, Int J Solids Struct, 46: 3771-3781.

Lu, Y. and Tan, Y. (2011) Mitigation of building responses to DDC impacts by soft
and stiff wave barriers, J Vib Control, 17 (2): 259-277.

Madheswaran, C.K., Natarajan, K., Sundaravadivelu, R. and Boominathan, A. (2009)


Effect of open or concrete-infilled trenches on screening of ground vibration
during pile driving, Exp Mech, 33(2): 43-51.

78
Massarsch, K.R. (2005) Vibration isolation using gas-filled cushions, International
Workshop on the Mitigation and Countermeasures of Ground Environment,
26-27 September, Shiga, 123-133.

McNeill, R.L., Margason, B.E. and Babcock, F.M. (1965) The Role of soil dynamics
in the design of stable test pads, Guidance and Control Conference, 16
August, Minneapolis, 366-375.

Miller, G.F. and Pursey, H. (1955) On the partition of energy between elastic waves
in a semi-infinite solid, Proc Roy Soc Lond Ser A, 233(1192): 55-69.

Murillo, C., Thorel, L. and Caicedo, B. (2009) Ground vibration isolation with
geofoam barriers: Centrifuge modeling, Geotext Geomembranes, 27: 423-
434.

Orehov, V.V., Moghanlou, R.N., Negahdar, H. and Varagh, A.M.B. (2012)


Investigation effects of trench barrier on the reducing energy of surface waves
in soils, 15th World Conference on Earthquake Engineering, 24-28
September, Lisboa, 10 pages.

Padade, A.H. and Mandal, J.N. (2012) Behavior of expanded polystyrene (EPS)
geofoam under triaxial loading conditions, EJGE, 17(S): 2543-2553.

Park, C.B., Miller, R.D. and Xia, J. (1999) Multichannel analysis of surface waves,
Geophysics, 64(3): 800-808.

Qiu, B. (2015) Numerical study on vibration isolation by wave barrier and


protection of existing tunnel under explosions, Ph.D. Thesis, Institut National
des Sciences Appliquées de Lyon, Lyon, 192 pages.

Qiu, B., Ali, L. and Irini, D.M. (2014) Numerical study of wave barrier and its
optimization design, Finite Elem Anal Des, 84: 1-13.

Rayleigh, J.S. (1885) On waves propagated along the plane surface of an elastic
solid, London Mathematical Society, 1 November, London, Oxford Journals,
17: 4-11.

Richart, F.E., Hall, J.R. and Woods, R.D. (1970) Vibrations of soils and foundations,
Prentice-Hall, New Jersey, 414 pages.

79
Saikia, A. (2014) Numerical study on screening of surface waves using a pair of
softer backfilled trenches, Soil Dyn Earthq Eng, 65: 206-213.

Saikia, A. and Das, U.K. (2014) Analysis and design of open trench barriers in
screening steady-state surface vibrations, Earthq Eng Eng Vib, 13: 545-554.

Sanchez-Sesma, F.J., Weaver, R.L., Kawase, H., Matsushima, S., Luzon, F. and
Campillo, M. (2011) Energy partitions among elastic waves for dynamic
surface loads in a semi-infinite solid, B Seismol Soc Am, 101(4): 1704-1709.

Shrivastava, R.K. and Kameswara Rao, N.S.V. (2002) Response of soil media due to
impulse loads and isolation using trenches, Soil Dyn Earthq Eng, 22: 695-
702.

Tsai, P.H. and Chang, T.S. (2009) Effects of open trench siding on vibration-
screening effectiveness using the two-dimensional boundary element method,
Soil Dyn Earthq Eng, 29: 865-873.

Tsai, P.H., Feng, Z.Y. and Jen, T.L. (2008) Three-dimensional analysis of the
screening effectiveness of hollow pile barriers for foundation-induced vertical
vibration, Comput Geotech, 35: 489-499.

Turan, A., Hafez, D. and El Naggar, M.H. (2013) The performance of inclined secant
micro-pile walls as active vibration barriers, Soil Dyn Earthq Eng, 55: 225-
232.

Wang, J.G., Sun, W. and Anand, S. (2009) Numerical investigation on active


isolation of ground shock by soft porous layers, J Sound Vib, 321: 492-509.

Wang, Z.L., Li, Y.C. and Wang, J.G. (2006) Numerical analysis of attenuation effect
of EPS geofoam on stress-waves in civil defense engineering, Geotext
Geomembranes, 24: 265-273.

Woods, R.D. (1968) Screening of surface waves in soils, Ph.D. Thesis, University of
Michigan, Detroit, 57 pages.

Xia. J., Miller, R.D. and Park, C.B. (1999) Estimation of near-surface shear-wave
velocity by inversion of Rayleigh waves, Geophysics, 64(3): 691-700.

80
Xu, B., Lu, J.F. and Wang, J.H. (2010) A semi-analytical model for the simulation of
the isolation of the vibration due to a harmonic load using pile rows
embedded in a saturated half-space, Open Civ Eng J, 4(1): 38-56.

Zoccali, P., Cantisani, G. and Loprencipe, G. (2015) Ground-vibrations induced by


trains: Filled trenches mitigation capacity and length influence, Constr Build
Mater, 74: 1-8.

81
APPENDIX

Appendix A. Results of Standard Penetration Tests

Table A.1. Standard penetration test log (Borehole 1)

Sample name Depth (m) (N60)' Description


SPT-1-15 1.5 40 Clayey sand
SPT-1-30 3.0 37 Clayey sand
SPT-1-45 4.5 24 Clayey sand
SPT-1-60 6.0 23 Soft clay
SPT-1-75 7.5 Refusal Stiff clay
SPT-1-90 9.0 21 Silty clay
SPT-1-105 10.5 Refusal Clay
SPT-1-120 12.0 Refusal Clay
SPT-1-135 13.5 Refusal Clay
SPT-1-150 15.0 Refusal Stiff clay
SPT-1-165 16.5 Refusal Stiff clay
SPT-1-180 18.0 Refusal Stiff clay
SPT-1-195 19.5 Refusal Stiff clay
SPT-1-210 21.0 Refusal Stiff clay
SPT-1-225 22.5 Refusal Stiff clay
SPT-1-240 24.0 Refusal Stiff clay
SPT-1-255 25.5 Refusal Stiff clay
SPT-1-270 27.0 Refusal Stiff clay
SPT-1-285 28.5 Refusal Stiff clay
SPT-1-300 30.0 Refusal Stiff clay

82
Table A.2. Standard penetration test log (Borehole 2)

Sample Name Depth (m) (N60)' Description


SPT-2-15 1.5 37 Silty clayey sand
SPT-2-30 3.0 32 Clayey sand
SPT-2-45 4.5 31 Clayey sand
SPT-2-60 6.0 29 Sandy clay
SPT-2-75 7.5 Refusal Stiff clay
SPT-2-90 9.0 Refusal Stiff clay
SPT-2-105 10.5 19 Clay
SPT-2-120 12.0 Refusal Stiff clay
SPT-2-135 13.5 Refusal Stiff clay
SPT-2-150 15.0 Refusal Stiff clay

Table A.3. Standard penetration test log (Borehole 3)

Sample Name Depth (m) (N60)' Description


SPT-3-15 1.5 37 Silty sand
SPT-3-30 3.0 31 Clayey sand
SPT-3-45 4.5 19 Clayey sand
SPT-3-60 6.0 Refusal Stiff clay
SPT-3-75 7.5 32 Clay
SPT-3-90 9.0 Refusal Stiff clay
SPT-3-105 10.5 Refusal Stiff clay
SPT-3-120 12.0 Refusal Stiff clay
SPT-3-135 13.5 Refusal Stiff clay
SPT-3-150 15.0 Refusal Stiff clay

83
Table A.4. Standard penetration test log (Borehole 4)

Sample Name Depth (m) (N60)' Description


SPT-4-15 1.5 40 Clayey sand
SPT-4-30 3.0 Refusal Stiff clayey sand
SPT-4-45 4.5 34 Clayey sand
SPT-4-60 6.0 28 Clay
SPT-4-75 7.5 28 Clay
SPT-4-90 9.0 Refusal Stiff clay

Table A.5. Standard penetration test log (Borehole 5)

Sample Name Depth (m) (N60)' Description


SPT-5-15 1.5 46 Silty sand
SPT-5-30 3.0 32 Clayey sand
SPT-5-45 4.5 33 Clayey sand
SPT-5-60 6.0 Refusal Stiff clay
SPT-5-75 7.5 Refusal Stiff clay
SPT-5-90 9.0 Refusal Stiff clay

84
Appendix B. Results of Sieve Analyses

Figure B.1. Grain size distribution of the soil sample BH-1-15 (Borehole 1, depth 1.5m)

85
Figure B.2. Grain size distribution of the soil sample BH-2-30 (Borehole 2, depth 3.0m)

86
Figure B.3. Grain size distribution of the soil sample BH-1-60 (Borehole 1, depth 6.0m)

87
Figure B.4. Grain size distribution of the soil sample BH-3-60 (Borehole 3, depth 6.0m)

88
Figure B.5. Grain size distribution of the soil sample BH-1-75 (Borehole 1, depth 7.5m)

89
Figure B.6. Grain size distribution of the soil sample BH-3-120 (Borehole 3, depth 12.0m)

90
Figure B.7. Grain size distribution of the soil sample BH-1-135 (Borehole 3, depth 13.5m)

91
Figure B.8. Grain size distribution of the soil sample BH-2-135 (Borehole 2, depth 13.5m)

92
Figure B.9. Grain size distribution of the soil sample BH-1-225 (Borehole 1, depth 22.5m)

93
CURRICULUM VITAE

Personal Information
Name, Surname :Onur Toygar
Nationality : Turkish (TC)
Date and Place of Birth : 07/12/1990, İstanbul/TURKEY
Maritial Status :Single
Phone : +90 532 718 3926
E-mail : onurtoygar@mu.edu.tr
onur.toygar@hotmail.com

Education
Degree Institution Year of Graduation
High School Antalya Private Science High School 2007
Muğla Sıtkı Koçman University
B.S. 2012
Department of Civil Engineering

Work Experience
Year Place Enrollment
2010 A Z Construction Company Trainee-Site Engineer
2011 A C Construction Company Trainee-Site Engineer
2012 Hande Building Audit Company Control Engineer
2012-Present Muğla Sıtkı Koçman University Research Assistant

Foreign Languages
English
Memberships
2012- Present TMMOB-İMO (Union of Chambers of Turkish Engineers and
Architects-Chamber of Civil Engineers)
2014- Present ZMGM (Association of Soil Mechanics and Geotechnical
Engineering)

94
Project Experience
Scientific Research Project
Ülgen, D., Toygar, O. (2013-2015), “Hendek Tipi Dalga Bariyerlerinin Titreşim
Yalıtımı Performansının Saha Deneyleri İle İncelenmesi”, Project Code:
13/05, Muğla Sıtkı Koçman University (as a researcher)
Publications
Article Published in SCI Journals
Ülgen, D. and Toygar, O. (2015) Screening Effectiveness of Open and In-Filled
Wave Barriers: A Full-Scale Experimental Study, Constr Build Mater, 86:12-
20.
Oral Presentations in International Symposium

Ülgen, D. and Toygar, O. (2014), Investigation on Vibration Isolation Performance


of Open Trench Barriers under Impact Loading, International Civil
Engineering Symposium for Academicians (ICESA), 17-20 May, Antalya,
Turkey, Geotechnical Engineering Volume:73-80.
Oral Presentations in National Symposium

Toygar, O. and Ülgen, D. (2015), Hendek Tipi Dalga Bariyerlerinin Titreşim


Yalıtımına Etkisinin Saha Deneyleriyle İncelenmesi, 6. Geoteknik
Sempozyumu, 26-27 November, Adana, Turkey, 8 pages.

95

You might also like