Sei sulla pagina 1di 17

Soil Dynamics and Earthquake Engineering 19 (2000) 371387

www.elsevier.com/locate/soildyn

Ground vibrations from sheetpile driving in urban environment:


measurements, analysis and effects on buildings and occupants
G.A. Athanasopoulos*, P.C. Pelekis
Department of Civil Engineering, University of Patras, GR-26500 Patras, Greece
Accepted 19 February 2000

Abstract
Following a comprehensive review of the subject of man-made ground vibrations, measurements of ground vibration caused by vibratory
sheetpile driving in recent soil deposits are reported in terms of particle velocities vs. distance from the source of vibration. The measurements were conducted on paved surfaces and sidewalks in the inner urban environment. Reconstructed particle displacement paths indicated,
predominantly, vertical vibrations of the Rayleigh type. The attenuation rate of vibrations with distance was compared to published results of
other studies and satisfactory agreement was found to exist. Values of particle velocity measured in this study, however, were lower than
corresponding values of other studies under comparable values of rated vibratory kinetic energy. This is possibly due to different soil
conditions. Average and upper bound linear loglog attenuation relationships are proposed, which t the results of measurements and are
representative of the conditions likely to be encountered in the urban environment. Measurement of vibrations on higher oors of multistory
reinforced concrete buildings indicated a signicant amplication of vertical vibration and an average curve for amplication magnitude vs.
oor level was tted to the results of measurements. A comparison of measured values of vibration with the observed performance of
buildings and with damage threshold values suggested by existing codes and standards indicated that the latter do not provide safety against
damage caused by vibratory densication of loose sandy soils. On the other hand, the existing criteria for human exposure to vibrations in
buildings, according to the results of this study, seem to adequately dene the degrees of human discomfort. q 2000 Published by Elsevier
Science Ltd.
Keywords: Man-made vibrations; Ground vibrations; Construction vibrations; Sheetpile driving; Amplication; Vibration criteria; Building damage; Human
response

1. Introduction
Ground vibrations can be generated either by natural
phenomena or by human activities. Among natural phenomena earthquakes (and the ocean waves) are the source of
ground vibrations of most interest. In the case of earthquakes the intensity of ground shaking may be high enough
to result in heavy structural damages or even collapse of
structures accompanied by loss of life. The study of all
aspects of earthquake occurrence and of the associated
effects on ground and soil structures is the subject of
geotechnical earthquake engineering [1].
Ground vibrations generated by human activities are
called man-made vibrations and vary greatly in intensity
depending on the particular source of vibration [2]. The
vibration intensity, however, is generally much lower
compared to the earthquake shaking intensity. The seismic
waves associated with man-made vibrations propagate in
* Corresponding author. Tel.: 1 30-61-997-677; fax: 1 30-61-997-274.
E-mail address: geolab.gaa@upatras.gr (G.A. Athanasopoulos).
0267-7261/00/$ - see front matter q 2000 Published by Elsevier Science Ltd.
PII: S 0267-726 1(00)00008-7

the ground and inevitably interact with above-ground or


in-ground engineered structures. This interaction induces
vibrations in the structure, which may disturb the people
occupying the structure and, in extreme cases, threaten its
serviceability and integrity. The study of man-made vibrations and their effects on structures constitutes a very interesting subject of soil dynamics and has received
considerable attention during the last two decades [3,4].
At present, a rather large number of codes, standards and
guidelines referring to the effects of groundborne man-made
vibrations on structures and humans are in existence. Their
provisions are used by designers, contractors and engineers
in an attempt to control and minimize the impact of vibrations on both the public and the wider environment.
However, due to the complexity of the phenomenon and
the great number of associated parameters the application
of pertinent specications is difcult and sometimes confusing. It is, therefore, generally recognized that a need exists
for obtaining more eld measurements of man-made vibrations and correlating the results to the observed behavior of
structures and people.

372

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

The current study presents a review of the state of knowledge on the subject of man-made, particularly pile-driving
vibrations, as well as the results of measurements of ground
vibrations generated by vibratory sheetpile driving in the
urban environment. Vibration levels were measured on
ground surface at increasing distances from the source and
on the higher oors of multistory reinforced concrete buildings in the vicinity of sheetpile driving. Measured values of
vibration intensity are correlated with observed effects on
the integrity of adjacent structures and with the provisions
of existing codes, standards and specications. The results
of measurements are also analyzed in terms of human
response, based on the reactions of the occupants of the
buildings (as well as of the authors themselves) and
compared with the provisions of pertinent standards.

2. Review of literature
2.1. Man-made ground vibrations
Peaceful human activities that generate ground vibrations
may be classied into the following main categories: (1)
operation of machines; (2) road and railway trafc; and
(3) construction activities [5]. Since the intensity of ground
shaking produced by man-made vibrations is much lower
compared to earthquake shaking these vibrations cannot, in
most cases, cause serious structural damage and their effects
are limited to the development of cosmetic cracks in the
walls and oors of building. Man-made vibrations can
also induce permanent deformations (densication) in
sandy soils, which are followed by foundation settlement.
This settlement has the potential of inducing more serious
structural damage than cosmetic cracking that results from
direct vibration. More likely than cosmetic cracking are
discomfort to the occupants of the affected buildings,
disruptions of occupant business activities and damage to
vibration-sensitive equipment. Human response to ambient
vibrations is an important aspect of man-made vibration
studies, and in many cases it is the disturbance and annoyance of people experiencing the vibration that triggers the
suspicion of structural damage and subsequent litigation
[3,6]. When man-made vibrations occur in the inner urban
environment the problems described above are intensied
due to the close proximity of buildings and other structures
to the source of vibrations. This explains why the study of
man-made ground vibrations constitutes a most important
and rapidly developing area of soil dynamics.
In man-made ground vibration studies it has been found
convenient to describe the intensity of shaking by the peak
value of particle velocity. The reason for this choice is the
well-established correlation between particle velocity and
observed cosmetic cracking [3,7], which is explained theoretically by the fact that the strain induced in the ground
during vibration is proportional to the particle velocity. The
particle velocity at a point of interest is usually measured in

three mutually perpendicular directions, one of them always


coinciding with the vertical direction (V-direction). From
the two horizontal directions, one coincides with the
sourcereceiver radial direction (R-direction) and the
other is perpendicular to the radial direction and is called
transverse direction (T-direction). The resultant intensity of
vibrations is expressed either as the maximum single value
of the three directional components (peak component) or by
the true vector sum of the three components (TVS). In some
circumstances the vibration intensity is simply described by
the peak vertical component of particle velocity. The square
root of the sum of the squares (SRSS) of maximum values of
the three components (irrespective of the time of occurrence) has also been used in the past. This latter approach
is now considered as overly conservative and should not be
used [3,7,8]. Field data seem to indicate that the peak
component particle velocity may be up to 25% lower than
the TVS value, whereas the SRSS value may exceed the
TVS value by 50% [3,8,9].
Ground vibrations from the operations of (heavy)
machines are mostly continuous and periodic and have
been studied extensively during the last 50 years by both
analytical and experimental methods [1013]. The focus
of the relevant studies has been on the development of
methods for estimating the magnitude of the expected vibrations from machine foundation. Fewer studies have focused
on the establishment of vibration criteria to protect the
integrity of the machine and avoid the disturbance of people
working or living in the immediate vicinity.
Ground vibrations caused by road and railway trafc are
in most cases random motions and have been the subject of
numerous studies in the past 25 years. These studies
involved eld measurements [1420] and they focused on
developing: (1) methods for empirical prediction of the
intensity of ground vibrations generated at a given distance
from typical vehicular trafc; and (2) numerical models encompassing the source of vibration and the transmission path for
prediction of vibration level at different distances from the
source and to the study of the effects of several parameters.
Construction vibrations have been the subject of intensive
study during the last few decades [3,4,79,2123]. Usual
construction activities encompass blasting operations in
rock sites, explosive demolition, pile driving in soils, vibratory or dynamic compaction of soils, deep excavation or
tunneling in either soil or rock sites and mechanized
construction activities in general. Pile and sheetpile driving
are probably the most frequent source of ground vibration
since in most of today's projects piles are used as foundation
elements whereas sheetpiles are employed as earth retainment elements and groundwater barriers. This explains why
pile-driving vibrations have caused much concern and are
covered by legislation in several countries.
2.2. Ground vibrations from pile driving
It has been known for a long time that pile driving creates

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

373

Fig. 1. Generation mechanism of seismic waves during vibratory (or impact) driving of piles in homegeneous soil (adapted from Woods [24]).

vibrations in the ground and that, occasionally, those


vibrations can damage structures or disturb people (or
other activities) in the vicinity of the pile driving [24].
Pile driving is achieved by impact or vibratory hammers
and both processes generate ground vibrations. The
mechanism of generation of seismic waves during pile driving is illustrated in Fig. 1. Vertically polarized body shear
waves are generated, as the pile is driven, along the skin
of the pile and their wave front propagates outward
having the shape of a cone, with, very shallow
anglepractically along a cylindrical surface. At the tip
of the pile, the displacement of an amount of soil volume,
generates both compressional P-waves and shear S-waves
which propagate outward from the tip on spherical waveforms. When the P- and S-waves encounter the ground
surface, part of their energy is converted to surface Rayleigh
waves (R-waves) and part is reected back into the ground.
The R-waves have both vertical and horizontal components
of motion and propagate along the surface of the ground
decaying in amplitude in proportion to the square root of
distance. Fig. 2 shows that surface waves generated by the
reections of body waves may be developed quite close to
the pile-driving point [3].
The attenuation with distance of ground vibrations caused

by pile driving is frequently modeled by a linear loglog


relationship, i.e.:
v kr 2m

where v is the peak particle velocity at ground surface (usually


in vertical direction); r is the distance from the vibration source
(usually the horizontal distance from the point of pile driving),
m is the slope or attenuation rate and k is a constant.
Results of measurements of ground vibrations generated
by vibratory sheetpile driving reported in the literature are
shown in Fig. 3. The lines shown are best-t lines to a
number of data points under rather hard driving conditions
[3,2527]. The results of Clough and Chameau [25] were
obtained by using an ICE 816 vibratory hammer with a rated
energy approximately equal to 5500 N m. Values of driving
energy for the rest of the data are not known. It may be seen
from the plot of Fig. 3 that the values of attenuation rate
(which is expected to depend on the ground type according
to Woods [24]) do not differ signicantly in the reported
cases. The vertical position of the best-t lines, however,
depends obviously on the amount of available driving
energy, on the type of pile and on soil conditions. In an
attempt to normalize and unify plots like the ones shown
in Fig. 3, Attewell and Farmer [28] have proposed the use of

374

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

Fig. 2. Determination of the minimum distance from the driving at which surface waves are generated (adapted from Dowding [3]).

scaled distance relationships in which, in addition to the


distance from the vibration source, the kinetic energy is also
taken into account (see also Moore et al. [29]). A simplied
relationship is given by Eq. (2):
v 0:75

p
E0
r

where v is the mean value of vertical peak particle velocity

(mm/s), r is the distance from vibration source (m) and E0 is


the kinetic energy (N m).
Eq. (2) has been subsequently modied by Attewell et al.
[30] in the form of quadratic loglog relationship to provide
a more satisfactory prediction of vertical ground vibrations,
as shown in Eq. (3):
p
p
E0
E0
2 0:234 log2
3
log v1 20:073 1 1:38 log
r
r

Fig. 3. Attenuation of vibrations with distance for vibratory sheetpile driving.

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

375

Table 1
Geometric attenuation coefcients for various sources of vibration (from
Kim and Lee [27])

Table 2
Proposed classication of earth materials by material attenuation coefcient
(from Woods [24])

Source location

Source type

Induced wave

Surface

Point

Body wave
Surface wave
Body wave
Surface wave

2.0
0.5
1
0

Class Material damping coefcient


a (1/m)

Body wave

1.0
0.5

Innite line
In-depth

Point
Innite line

where v1 is the value of peak particle velocity one standard


deviation above the mean value, r is the distance from vibration source (m) and E0 is the kinetic energy (N m).
Values of particle velocities predicted by Eq. (3), for
kinetic energies ranging from 900 to 5000 N m, have been
plotted in the diagram of Fig. 3. It is observed that the level
of vibrations predicted by Eq. (3) (which is actually an
upper bound or conservative value) does not differ substantially from the measured values, which were mean values
obtained under hard driving conditions.
The attenuation of ground vibration amplitude with
distance can be alternatively described by the Bornitz equation [10]:
 n
r
4
w2 w1 1 e2ar2 2r1
r2
where w1 and w2 are vibration amplitudes at distances r1 and
r2 from the source of vibration, n is the geometric damping
coefcient and a is the material damping coefcient.
The value of geometric damping coefcient, n, depends
on the type of seismic waves and takes the values indicated
in Table 1 [27], for surface and in-depth point or innitely long sources. It may be seen that for surface waves
the geometric decay of amplitude is proportional to the
square root of distance, whereas for body waves the decay
is proportional to the rst power of distance. For body waves
traveling at the ground surface, the decay is particularly
intense and is proportional to the square of the distance.
Attenuation due to material damping in Eq. (4) is
described by the coefcient a, which depends on soil type
and increases linearly with frequency of vibration. Values of
a for four classes of soils and two bounding values of
frequency were reported by Woods and Jedele [31] and
are summarized in Table 2. Yang [32] has recently proposed
the use of an energy attenuation coefcient a0 which is
independent of frequency and is related to the conventional
coefcient described above by the equation a0 a=f ; where
f is the frequency of vibration. Values of a0 coefcient
proposed by Yang [32] for different types of soils are
summarized in Table 3.
Attenuation relationships in the form of Eq. (4) have the
advantage of separating the effects of geometric and material damping. They can be implemented, however, only in
cases where the vibration amplitude is known at a reference

5 Hz
I
II

0.010.03
0.0030.01

III
IV

0.00030.003
, 0.0003

Description of material

50 Hz
0.10.3
0.030.1

Weak or soft soils NSPT , 5


Competent soils 5 , NSPT ,
15
0.0030.03 Hard soils 15 , NSPT , 50
, 0.003
Hard, competent rock NSPT .
50

point at a known distance from the source (e.g. when


preconstruction tests have been performed) and when the
type of propagating waves can be predicted with certainty.
Woods [24] concludes that the pseudoattenuation approach
implemented in Eqs. (1)(3) is satisfactory for the short
distance range and, therefore, is appropriate for pile-driving
vibrations for which the concern is focused in the near eld.
The empirical relationships described by Eqs. (1)(4), are
of considerable value to piling practitioner, being easy to
apply for dened hammer and known soil conditions.
However, they do not account for strongly non-uniform
soils or allow investigation of the types of seismic wave
generated and of dynamic soilstructure interaction. Some
progress has been made, recently, in these directions by
developing computational (nite element) models that
incorporate the source of vibration and the near- and fareld transmission path as shown in Fig. 4 (Ramshaw et al.
[33,34]). Potential applications of such computational
modeling include assessments for the effects of different
pile-driving methods, non-uniform soil conditions and
wave interaction with buried services and typical structures.
Another approach to overcome the problem of unknown
or uncertain soil conditions and unknown soilstructure
interaction mechanism is the application of the impulse
response function concept, proposed by Svinkin [35]. In
this approach an experimental impulse response function
is obtained in the eld, before the start of actual operations,
by using a drop weight as a source of vibration and recording the response at points of the ground and of the affected
structures as shown in Fig. 5. The expected vibration levels
at the points of interest are then computed in terms of the
actually applied loading from the particular vibration
source. The rst applications of the method show satisfactory correlation between predicted and measured vibrations.
2.3. Allowable values of groundborne vibrationscodes
and standards
The ultimate goal of most studies on the subject of manmade ground vibrations is the development of methods for
assessing the effects of these vibrations on soil materials,
structures and on people living or working in the structures.

376

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

Fig. 4. Finite element modeling of seismic waves generated during pile driving (from Ramshaw et al. [34]).

Numerous specications, guidelines, regulations and code


provisions (at the international, regional and national level)
have been issued during the last 20 years by various agencies and institutions that dene allowable (or permissible)
values of man-made ground vibrations. These allowable
vibration values are usually given in terms of peak particle
velocity and are actually limiting values above which cracking of structures may be developed. Three different levels of
cracking may be categorized according to Dowding [3]: (1)
cosmetic cracking, i.e. threshold damage such as opening of
old cracks and formation of new plaster cracks and dislodging of loose structural particles (e.g. loose bricks from
chimneys); (2) architectural cracking or minor damage
(i.e. supercial damage, not affecting the strength of the
structure, such as broken windows, loosened or fallen plaster and hairline cracks in masonry); and (3) structural cracking or major damage that results in serious weakening of the
building (large cracks, shifting of foundations or bearing
walls, major settlement resulting in distortion or weakening
of the structure, wall put out of plump).
Skipp, [36] has recently presented a brief review of existing standards for groundborne man-made vibrations. In
most standards the intensity of vibrations is described in
terms of peak particle velocity (usually peak component
velocity) measured at ground level at a point close to the
structure, and permissible (or no-cosmetic cracking) values
are given for different ranges of dominant frequency. The
allowable values range, in general, from 3 to 70 mm/s with
the lower values pertaining to old and small residential
buildings and the higher values to modern large size
commercial or industrial buildings. The corresponding
frequency range is 1060 Hz. Other codes and standards
are based on diagrams that provide the threshold values as
a function of frequency of vibration (frequency range from 1
to 100 Hz). In Fig. 6 the provisions of four frequently used
national codes, i.e. U.S. Ofce of Surface Mining (OSM) as
reported in Siskind et al. [37], German Institute of Standards
(DIN), [38], British Standards Institution (BSI) [39,40] and
the Swiss Association of Highway Engineers (SN) [41] are,
summarized. It may be seen from the plots of Fig. 6 that the

threshold values increase with the frequency of vibration


and depend on the type and the construction quality of the
building. It should be noted that the low allowable limits
specied by DIN and SN are not based on scientic observation of cracking but they are, rather, administrative guidelines to control annoyance. Dowding [3], stresses the lack of
data for the foundation of the German DIN 4150 standard;
he also emphasizes the fact that the data on which the OSM
standard was based, have included observation of 150-yearold very sensitive homes that did not suffer cosmetic cracking for particle velocities up to 20 mm/s.
Massarsch and Broms [42] have proposed an interesting
approach for estimating the threshold values of particle
velocity by using a simple relationship incorporating the
shear wave propagation velocity in the ground and coefcients to account for vibration source, building category and
degree of damage. Comparison of the results obtained by
using this approach with the provisions of a number of codes
indicated a very good correlation.
The limiting values of particle velocity summarized in
Fig. 6 correspond to observed direct damages from ground
vibrations. In the case of pile driving such direct effects are
Table 3
Values of energy attenuation coefcient, a 0, for various soil and rock types
(from Yang [32])
a0 ( 10 23 s/m)

Soil group
Rocks (covering layer within
1.52.0 m)
Hard plastic clays
Broke stones of medium density
cobbles
Plastic clays, coarse sands and
gravels of medium density
Soft plastic clays, silts, slightly
dense, medium or coarse sands
Silty clays, silts and saturated
ne sands
Recently deposited clays and
unsaturated loose sands

Shale, limestone

0.3850.485

Sandstone

0.5800.775
0.3850.525
0.8501.100
0.9651.200
1.2551.450
1.2001.300
1.8002.050

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

377

Fig. 5. Experimental determination of impulse response functions (from Svinkin [35]).

usually limited to within one pile length distance from the


point of pile installation [3,24]. More serious structural
damage can be induced, however, in the case of loose
sandy or silty soils from vibration-induced settlement of
the ground. Such settlement-induced damages, especially
when a large number of piles are being driven, may occur
as far as 400 m from the source of vibration [24]. Several
cases have been reported in the literature, in which signicant settlement occurred as a result of pile or sheetpile driving [25,26,4346]. In some of these cases settlements
occurred under vibration levels v 2 mm=s well below
the allowable levels usually specied by the codes and
specications presented above, which are based on the
direct effects of vibrations on structures. There is, then,
clearly a need for improvement of the existing vibration
standards and for development of settlement criteria for
soils susceptible to vibratory densication. Some recent
research is focusing in developing methods for predicting
vibration-induced settlements in sandy soils [4751]

whereas Woods [24] concludes that simple methods of estimating the magnitude of settlement are still out of reach.
A number of standards and guidelines exist that specify
threshold values of vibrations with regard to different
aspects of human sensitivity. These threshold values are
functions of the frequency and the duration of vibrations
and may be expressed in terms of peak values of acceleration [52], velocity [21] or displacement [4] of vibrations. In
Fig. 7 a number of provisions and recommendations regarding limiting values of continuous vibration intensity for
different levels of human discomfort (including the BSI
[53] and ANSI [54] standards) are summarized. It should
be noted that the provisions of the International Standards
Organization (ISO) are identical to the ANSI standards as
reported in Woods [24]. It may be observed in Fig. 7 that the
threshold valuesin terms of particle velocity and displacementare decreasing with increasing values of the
frequency of vibration. It is assumed that the vibration
intensity indicated in the plots of Fig. 7 has been measured

Fig. 6. Comparison of various threshold vibration criteria for structural damage.

378

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

Fig. 7. Threshold vibration criteria for various degrees of human discomfort in terms of: (a) acceleration; (b) velocity; and (c) displacement of vibrations.

on the surface of building oors where people are standing.


Kelley et al. [55] emphasize the fact that ground vibration
criteria (i.e. vibrations measured at the base of a multistory
building) alone are not sufcient to predict occupant
discomfort conditions, since the individual building
response can result in amplied vibration magnitude on
elevated oors. In this regard studies of amplication of
base vibration could provide valuable data in order to relate
the conventional construction vibration limits to human
discomfort criteria at upper oors of buildings.
3. Results of eld measurements
The results of ground vibration measurements reported in
this study were obtained during sheetpile driving at several
sites of the central part of the city of Patras, Greece. The
construction project (which is still in progress) was undertaken by Patras Water and Sewer Municipal Enterprise and
involved the installation of a new, 1400 mm diameter, interceptor sanitary sewer at depths ranging from 3.5 to 7.5 m
under the surface of the ground. The route of the sewer
crosses densely populated areas of the city and follows the
direction of some major (and few minor) streets and roadways of the urban complex. Temporary support of the sides
of trenches excavated for the installation of the sewer was
achieved by braced sheetpile walls, whereas the driving of
the sheet piles was accomplished by vibratory hammers.
Sheetpile driving in an urban environment and at short
distances from residential and commercial buildings raised
concerns regarding the effects of groundborne vibrations on
the buildings and their occupants. It was, therefore, decided
to implement a plan for monitoring ground vibrations during
the sheetpile-driving operations in an attempt to avoid

actual or alleged structural damage. The monitoring


program was commissioned to the authors of the paper
who conducted vibration measurements at varying distances
from the point of sheetpile installation on the surface of
street pavements and sidewalks as well as on the ground
oor of adjacent buildings. Frequent complaints of the occupants of adjacent multistory buildings also made it necessary to conduct vibration measurements at the higher oors
of the buildings. The implementation of the vibration monitoring program resulted in an extensive set of data on
groundborne vibrations from vibratory sheetpile driving
which are presented and analyzed in the current study.
3.1. Instrumentation and equipment used for vibration
measurements
The vibration measurements reported in this study were
conducted using vertical and horizontal geophones with a
natural frequency equal to 2 Hz. The measurements at each

Fig. 8. The system used for measuring particle velocities in the present
study.

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

379

Table 4
Details of vibration measurements at sites of sheetpile driving
Site Soil conditions

Fill (03.0 m), sandy


silt (3.08.0 m) and
clay of low plasticity
(8.016 m). (water
table at 21.0 m)
Clayey sand (03.5 m),
sandy silt (3.5
16.0 m). (water table at
23.0 m)
Clay (01.5 m), sand
with some silt, clay and
gravel (1.54.0 m),
silty sand (4.08.0 m).
(water table at 23.0 m)
Sandy clay, sandy silt
(08.0 m), silty sand
(8.0). (water table at
23.0 m)
Fill (0.01.0 m), silty
sand-gravel (1.0
3.0 m), sandy silt (3.0
4.5 m), clay of low
plasticity (4.5-). (water
table at 21.0 m)
Sandy clay, sandy silt
(08.0 m), silty sand
(8.0). (water table at
23.0 m)
Clayey ll (02.0 m),
gravel (2.03.5 m),
silty sand with gravel
(3.55.5 m), clay of
low plasticity (5.5).
(water table at 22.0 m)
Similar to Site F

Similar to Site F

Pile type

Types of
vibratory
hammer

Frequency
(Hz)

MGF RBH
60M

Peak component particle velocity, (mm/s)


Close to source

Outside of
building

Ground oor of
building

Upper oors
of buildings

13

7.0
(dist < 0.20 m)
(pavement)

0.45
(dist < 21.0 m)
(sidewalk)

0.70
(dist < 30.0 m)

1.8
(3rd oor)

ARBED PU16 MS-5H4


(length 10 m)

23

5.3
(dist < 1.50 m)
(pavement)

4.0
(dist < 6.80 m)
(sidewalk)

1.0
(dist < 10.0 m)

LARSSEN III
neu
(length 8 m)

MGF RBH
60M

20

15.0
(dist < 1.60 m)
(pavement)

5.5
(dist < 3.60 m)
(sidewalk)

1.8
(dist < 7.50 m)

LARSSEN 22
(length 8 m)

ABI RE 10000/ 40
3

25.0
(dist < 1.50 m)
(pavement)

2.2
(dist < 5.50 m)
(sidewalk)

4.0
(dist < 8.50 m)

7.0
(1st oor)

LARSSEN III
neu
(length 8 m)

MGF RBH
60M

25

10.0
(dist < 1.80 m)
(pavement)

7.5
(dist < 3.40 m)
(sidewalk)

1.4 (dist < 7.0 m)

LARSSEN III
neu
(length 7 m)

ICE 416

17

6.0
(dist < 5.15 m)
(sidewalk)

6.0
(dist < 3.60 m)
(sidewalk)

1.4 (dist < 9.0 m)

LARSSEN III
neu
(length 8 m)

MGF RBH
60M

21

5.0
(dist < 3.90 m)
(pavement)

20.0
(dist < 4.40 m)
(sidewalk)

2.0
(dist < 5.50 m)

2.0
(1st oor),
10.0
(4th oor)
-

LARSSEN III
neu
(length 8 m)

ICE 416

16

6.05
(dist < 2.40 m)
(pavement)

1.25
(dist < 11.40 m)
(sidewalk)

0.51
(dist < 13.40 m)

LARSSEN III
neu
(length 8 m)

ABI RE 10000/ 24
3

35.0
(dist < 1.0 m)
(pavement)

4.1 (dist < 7.0 m)

LARSSEN III
neu
(length 7 m)

point were conducted by using of a set of three geophones


positioned in such a way as to measure particle velocities in
the V-, R- and T-directions. The geophones were connected
to a portable PC Notebook equipped with a data acquisition
card and loaded with software appropriate for digital signal
acquiring and processing. A photograph of the system is
shown in Fig. 8. Time histories of particle velocity were
obtained at each point of measurements from which the
corresponding Fourier spectra were determined as well as

0.64
(1st oor),
1.10
(5th oor),
2.85
(7th oor)
2.05
(2nd oor),
1.55
(4th oor),
2.40
(6th oor)

the particle displacement paths on the radial and transverse


vertical planes. The results of all measurements were saved
on 3.5 in. diskettes for keeping a permanent record of data
and for subsequent further processing.
3.2. Results of measurements
The results of vibration measurements presented in the
current study were obtained at nine sites in the city of Patras

380

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

Fig. 9. Particle displacement paths during vibration at Site H (2.40 m from point of pile driving).

during vibratory sheetpile driving. Four types of vibratory


hammers with rated driving energy ranging from 1000 to
3000 N m were used for driving the two types of sheetpile
used for the braced sheetpile wall. The results of measurements for the nine sites, expressed as peak component particle velocities at selected locations, are summarized in Table
4 which also includes information regarding the soil prole
at each site, the length of the pile and the frequency of
recorded vibrations. Column 6 of Table 4, gives the value
of peak particle velocity recorded near the point of sheetpile
driving, whereas the level of vibration at a point outside the
affected building is given in column 7. At almost all sites the
intensity of vibrations was also recorded at the ground oor
of the building (column 8) whereas the values of peak vertical component of vibration at midspans of higher-level
oors are given in column 9. By comparing the peak component particle velocities along the three orthogonal directions, which were recorded at the nine sites, it was found
that in almost all cases the maximum value corresponded to
the vertical component of vibration. It should also, be noted
that increased values of particle velocities were observed
momentarily during run-up and shutdown of the vibratory

driving process of the sheet piles. According to Dowding [3]


this type of increase of ground vibration may be on the order
of fourfold and it occurs when the driving frequency passes
momentarily through the frequency that produces the maximum ground motion. In this study the particle velocities
listed in Table 4 are the steady-state values.
The soil conditions at the nine sites do not differ appreciably from each other, and consist of mixtures of silt, sand,
some gravel and clay up to a depth of 8.0 m from the ground
surface. Below that depth, in most of the sites, a layer of low
plasticity clay is encountered. The water table was encountered at depths ranging from 1.0 to 3.0 m from ground
surface.
3.3. Analysis of the test data
When ground vibration data along three orthogonal directions are available, it is possible to plot the particle displacement paths at each point of measurement and obtain
valuable information regarding the types of wave propagating away from the source of vibration. Such typical particle
displacement paths for this study are shown in Figs. 9 and

Fig. 10. Particle displacement paths during vibration at Site H (11.350 m from point of pile driving).

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

381

reections and wave interference and result in complicated


time history records.
The results of all vibration measurements are plotted in
Fig. 11 in the form of a peak component particle velocity vs.
distance attenuation diagram. A linear loglog best-t line
has been obtained by regression of the data points, which is
described by Eq. (5):
v 32r 21:5

where v is the mean value of peak particle velocity (mm/s)


and r is the distance from pile (m).
It is interesting to note that the value of the attenuation
rate, m, in Eq. (5) is m 1:5; in agreement with the value
proposed by Woods [24] for Class II soils. An upper bound
line is also shown in the diagram of Fig. 11, which is
described by Eq. (6):
vub 80r 21:5

Fig. 11. Results of ground vibration measurements obtained in the present


study during vibratory sheetpile driving and best-t mean and upper bound
lines.

10. Those displacements were obtained for Site H by


combining the vertical and horizontal time histories of
components of motion at two different distances from the
source of vibration (2.40 and 11.35 m). The required values
of displacements were obtained by integrating the corresponding velocity time histories. It may be seen that at a
short distance from the source of vibration the particle
motion occurs mainly in the radial vertical plane (very
small magnitude of displacement in the transverse vertical
plane). The corresponding displacement path, in this case,
has an elliptical shape similar to the R-wave motion, i.e. the
vertical component of motion is greater than the horizontal
one. At a greater distance from the source the motion
becomes predominantly vertical, i.e. the horizontal component of motion is further reduced whereas the motion of the
transverse vertical plane is slightly increased. The plots of
Figs. 9 and 10 indicate that sheetpile driving generates
waves that have the characteristics of R-waves; this observation is important when considering that in the urban environment the seismic waves propagate in a soil medium that
incorporates buried objects and is covered by pavements
and sidewalk structures which could generate multiple

where vub is the upper bound value of peak component


particle velocity in mm/s vub < 2:50v: Upper bound estimates of peak particle velocity may be used in applications
of sheetpile driving where no provision for monitoring the
ground vibrations has been made.
It was mentioned in Section 2.2 that Eq. (4), which separates the effect of geometric damping from the effect of
material damping, can also describe the attenuation of
ground vibration with distance. Such representations are
useful when analyzing vibration data from a single source
of vibration under homogeneous soil conditions at increasing distances from the source. In such cases it is possible to
nd the best-t values of coefcients n and a and draw
conclusions regarding the predominant types of wave propagated from the source and the amount of material damping
of soil materials of the site [27]. In the urban environment
where the underground conditions are far from homogeneous and the measurements are taken on paved surfaces
at short distances from the source, it is possible to observe
vibration intensities that are locally larger despite being
slightly further from the source due to reections and
constructive interference of waves. This local observation
does not change the physics that amplitudes decrease with
increasing distance in general.
In Fig. 12 the best-t relations described by Eqs. (5) and
(6) are compared to the data obtained by other investigators,
which were also presented in Fig. 3. It may be seen that the
value of attenuation rate of vibration intensity with distance
derived in the present study is approximately equal to the
attenuation rate of the rest of the data included in Fig. 12, for
distances up to 20 m from the source. The vertical position
of the mean and upper bound curves of the present study,
however, is lower than corresponding values of other
studies. In the present study the values of rated kinetic energies, E0, ranged from 1000 to 3000 N m, and the driving was
performed in relatively soft soil materials. It is then reasonable to expect lower levels of vibration compared to the

382

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

Fig. 12. Comparison between the results of this study and previous data presented in Fig. 3.

values reported by Clough and Chameau [25] (which were


obtained for E0 < 5500 N m and by Linehan et al. [26]
(which were obtained for E0 < 5000 N m: The value of
kinetic energy for the sheetpile-driving data reported by
Dowding [3] is not known; it is reported, however, that
the data pertain to very hard driving conditions. The data
reported by Kim and Lee [27], do not include information
on the driving energy and seem to indicate increased rates of
attenuation of vibrations at larger distances from the source.
In Fig. 12, the upper bound values of particle velocities
predicted by the Attewell et al. [30] and empirical equation
(Eq. 3) (with appropriate values of kinetic energy) are
higher than almost all other data included in the plot.
The above observations lead to the conclusion that for
relatively short distances from sheetpile driving in the
urban environment the rate of attenuation of vibration
with distance does not vary considerably for different
cases and the value of m 1:5 may be used for initial estimates. The intensity of vibrations, however, may vary
considerably (up to ve times) depending on the kinetic
energy delivered to the sheetpile during driving. The value
of the rated energy of the vibratory hammer can help,
however, in preliminary estimates of the expected intensity
of motion.
3.4. Effects of vibrations on structures and people
The ground vibration data obtained in the course of the
present study provide an opportunity to examine the measured
vibration intensities in the framework of the provisions of

pertinent standards and specications and compare the


expected vs. observed performance of structures and the
response of people. To facilitate the comparisons appropriate vibration data for the nine sites from Table 4, were
superposed to the vibration criteria diagram presented in
Section 2.3, as shown in Fig. 13. Peak component particle
velocities recorded at ground-level points just outside the
examined buildings are plotted in Fig. 13 vs. the measured
values of dominant frequency of vibration. It may be
observed that in all cases the vibration intensity was lower
or equal to the damage threshold values for modern
buildings.
In Site G the measured particle velocity was equal to
20 mm/s, a value that coincides with the USA OSM limiting
value but is lower than the BS 7385 value. At this site some
architectural cracking of the brick walls of a three-story
reinforced concrete building as well as a settlement and
cracking of the sidewalk adjacent to the building were
observed. The observed settlement of the site indicates
that the wall cracking was not induced by direct vibrations
but rather by vibratory densication of the layer of granular
soil materials that are encountered in this site between
depths of 2.0 and 5.5 m from ground surface.
It may, then, be concluded that use of the vibration criteria
summarized in Fig. 13, for cracking due to vibration only,
would have incorrectly indicated that the vibration levels
developed during sheetpile driving would not affect the
integrity of adjacent buildings. The observed performance,
however, indicates that building damage due to settlement
may be caused by vibration levels that are below the

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

383

Fig. 13. Superposition of vibration data obtained in this study on the diagrams of vibration criteria of Fig. 6.

limiting values of common standards and specications.


Thus the possibility of vibratory densication must always
be investigated as an issue separate from direct vibratory
cracking.
To assess the reliability of the criteria of human sensitivity to vibrations, the results of measurements have been
superposed, in Fig. 14, on the vibration criteria diagrams,
which are presented in Section 2.3. It may be seen that

according to the three types of criteria (acceleration, velocity, displacement vs. frequency) the level of vibrations at
the upper oors of the buildings was such that they could be
characterized as unpleasant, disturbing and possibly intolerable or painful. The authors of this paper have themselves
experienced these vibrations and are in a position to conrm
the validity and reliability of the threshold values depicted
in Fig. 14. The reliability of these threshold values was also

Fig. 14. Superposition of oor vibration data obtained in this study on the diagrams of vibration criteria for human discomfort of Fig. 7.

384

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

Fig. 15. Time histories of vertical vibrations and corresponding Fourier spectra recorded on the ground oor and higher oors of residential buildings at Sites A
and F during vibratory sheetpile driving.

veried by the numerous complaints expressed by the occupants of buildings adjacent to the locations of sheetpile
driving.
3.5. Vibration intensity at the higher oors of multistory
buildings
Very limited data are presently available regarding the
amplication (or deamplication) of man-made ground
vibrations at the higher oors of multistory buildings. This
is a critical issue, however, when studying the effects of
man-made vibrations, since it is the level of the oor vibration and not of the ground surface that should be checked
against the threshold levels dened by human sensitivity
standards and regulations.
Oriard et al. [56] reported measured values of upper oor
vibrations in high-rise (up to 22 oors) buildings in Atlanta,
USA, induced by nearby blasting, and compared them with
the measured base input motion. They found that in most
cases the level of vibration decreased in the upper oors of
the buildings. In some cases, however, the intensity of vibration in the upper oors was higher compared to the one at
intermediate oors. On the other hand, Kelley et al. [55],
reported case histories where man-made vibrations, induced
by trenching of slurry-wall panels and vibratory sheetpile
driving, resulted in amplication of base motions in two old
multistory (up to 11 oors) buildings with steel frames, in
Boston, USA.
Obviously, the response of building elements (e.g. oors,
walls) to man-made base vibrations depends on many parameters (dimensions, construction materials) and it is not

easy to establish general rules before obtaining a large


number of eld measurements. In the present study
measurements of vertical vibrations induced by sheetpile
driving, at the midspan of oors at different levels of typical
multistory (up to 7 oors) reinforced concrete buildings in
Greece, with an average span size of 4 m, were conducted,
as indicated in Table 4. Fig. 15 shows examples of time
histories and corresponding Fourier spectra of vertical particle velocities measured at the ground oor and higher oors
in reinforced concrete buildings at Sites A and E. The results
of all measurements are plotted in Fig. 16, as normalized
peak particle velocity (i.e. velocity measured at the particular oor divided by velocity measured at ground level) vs.
the level of the oor. For the normalization of higher oor

Fig. 16. Amplication of vertical particle velocity (relatively to input


motion) at the elevated oors of residential reinforced concrete buildings,
during nearby vibratory sheetpile driving.

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

velocities, values of velocity measured either outside or at


the ground oor of building could be used. Measurements
made outside the building have the advantage of being free
from the effects of structural details of the particular building on the transfer of energy from the outside to inside. For
this reason use of the outside values of velocity could
provide more reliable data for studying the amplication
of vibration along the height of the building. Table 4
includes results of measured ground-level vibration at
both outside and at the ground oor of the studied buildings.
Use of values of vertical motion obtained outside the buildings in this study (for normalization purposes) resulted in a
rather irregular pattern of amplication ratio (ranging from
0.33 to 4) along the height of the buildings. This irregular
pattern may be due to a peculiar vibration response of sidewalks and invalidates any attempt to derive vibration amplication values based on outside motion. It was thus decided
to use the ground oor motion for normalization purposes
despite the fact that the values of amplication ratio might
have been overestimated.
On the basis of ground oor motions the plot of Fig. 16
indicates an amplication of the vertical vibration of the
elevated oors in buildings, which is obviously caused by
the excitation of the natural frequency of the buildings
oors. Those natural frequencies, according to the Fourier
spectra of Fig. 15, are in the range of 1320 Hz in agreement with values suggested in the literature [3]. It may be
seen that the ground-level motion was amplied in all cases
at the upper oors of the buildings (with the exception of
Site I) and that the amplication ratio increased with the
level of the oor (up to 7). The rate of increase of amplication ratio decreased, however, with the level of the oor and
it should be anticipated that for high-rise buildings, the
motion at the upper oors would ultimately be deamplied.
The data for the building in Site I indicate a slight deamplication of vertical vibrations at the upper oors. This
differentiation of behavior cannot be easily explained; one
possible reason being the smaller span dimensions of the
particular oors on which the measurements were
conducted (2.5 to 3.0 m) compared to the dimension of
the spans of the rest of buildings.
Plots like the one shown in Fig. 16 provide valuable
information, at least for the particular types of examined
buildings. This information can be utilized for predicting
the expected higher oor vibration levels in terms of the
ground vibrations induced by sheetpile driving and for
establishing limiting values pertaining to the comfort of
the occupants of the buildings and undisrupted operation
of sensitive equipment.
4. Conclusions
Field measurements of ground vibrations produced by
vibratory sheetpile driving were made in an urban environment and provided values of particle velocities on paved

385

ground surfaces at varying distances (up to 30 m) from the


source of vibration and on the ground oor and higher oors
of adjacent buildings. The rated energy of the vibratory
hammer used for pile driving was on the average equal to
2000 N m, the penetration depth reached 10 m from the
ground surface and the frequency of vibration ranged from
13 to 40 Hz. The soil conditions consisted of alluvial deposits (gravel sands, silts and clays) with the water table
ranging from 1.0 to 3.0 m below the ground surface.
Based on the results of measurements the following
conclusions can be drawn:
1. The particle displacement paths indicated a predominantly vertically polarized Rayleigh wave with a small
amplitude motion in the radial direction.
2. The value of pseudoattenuation coefcient (i.e. the power
to which the distance is raised in a loglog linear particle
velocity vs. distance relationship) was found to be equal
to 21.5, a value equal to the one proposed by other
investigators for competent soils.
3. Vibration intensities measured in the present study, when
plotted against distance, were in general lower than
values reported in the literature and values predicted by
published empirical relationships. This difference may be
the result of different soil conditions.
4. The measured intensities of ground vibrations obtained in
this study were compared to the limiting values
suggested by relevant codes and regulations. It was
found that for soils susceptible to vibratory densication
(loose sandy soils) the threshold values for direct vibration damage cannot provide adequate safety against
building damages induced by foundation settlements.
5. Vertical vibration measurements at the ground oor (as
well as at ground points outside and adjacent to building)
and at higher oors of multistory residential reinforced
concrete buildings indicated that the intensity of vibrations is amplied signicantly along the height of the
building. It is those amplied values of particle velocities
that should be compared to the allowable values for
human exposure to vibrations provided by the relevant
codes.
6. Existing standards for the threshold values of human
sensitivity to vibration, based on the results of this
study, seem to classify adequately the degree of human
discomfort and should be used with condence for vibration control.

Acknowledgements
The authors express their thanks to the Patras Water and
Sewer Municipal Enterprise for permitting the publication
of the results of vibration measurements reported in this
study. Thanks are also due to eld supervising engineers
and construction workers of the construction rms involved
in the installation of the interceptor sewer.

386

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

References
[1] Kramer SL. Geotechnical earthquake engineering. Upper Saddle
River, NJ: Prentice Hall, 1996 (653pp.).
[2] Athanasopoulos GA, Pelekis PC. Man-made ground vibrations:
results of eld measurements. In: Theocaris PS, Fotiadis DI, Massalas
CV, editors. Proceedings of the Fifth National Congress on
Mechanics, vol. 1. The University of Ioannina, Greece, 1998. p.
21926.
[3] Dowding CH. Construction vibrations. Upper Saddle River, NJ:
Prentice Hall, 1996.
[4] Skipp BO. Ground vibration-codes and standards. In: Proceedings of
the Conference on Ground Dynamics and Man-made Processes. Institution of Civil Engineers, London, 1997.
[5] Massarsch KR, Madshus C, Bodare A. Engineering Vibrations and
solutions. In: Prakash S, editor. Proceedings of the Third International
Conference on Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics, vol. III. St. Louis, MO, 1995. p. 134953.
[6] Gubbe LW. Bad vibrations. Civil Engineering, ASCE. 1996:5860
(August).
[7] New BM. Construction induced vibrations in urban environment. In:
O'Rourke TD, Hobelman AG, editors. Excavation and support for the
urban infrastructure. Geotechnical Special Publication No. 33,
1992:21239.
[8] Mayne PW. Ground vibrations during dynamic compaction. In: Gazetas G, Selig ET, editors. Vibration problems in geotechnical engineering. Special Publication of ASCE, 1985:24765.
[9] Hiller DM, Hope VS. Groundborne vibration generated by mechanized construction activities. In: Geotechnical engineering. Institution
of Civil Engineers, 1998;131:22332.
[10] Richart Jr. FE, Hall Jr. JR, Woods RD. Vibrations of soils and foundations. Englewood Cliffs, NJ: Prentice Hall, 1970 [414pp.].
[11] Prakash S, Puri VK. Foundations for machines: analysis and design.
Canada: Wiley, 1988 (656pp.).
[12] Gazetas G. Foundation vibrations. Foundation engineering handbook,
New York: Chapman & Hall, 1990. p. 55393 [chap. 15].
[13] Kaynia AM, Mudshus C, Furuhorde R, Jostad HP. Impedance models
for machine foundation analyses. In: Proceedings of the Conference
on Ground Dynamics and Man-made Processes. Institution of Civil
Engineers, London, November 1997.
[14] Barneich JA. Vehicle induced ground motion. In: Gazetas G, Selig
ET, editors. Vibration problems in geotechnical engineering. Special
Publication of ASCE, 1985:187202.
[15] Al-Hunaidi MO, Rainer JH, Pernica G, Tremblay M. Trafc-induced
vibration in buildingsuse of site cut-off frequency as a remedial
measure. In: Cakmak AS, Brebbia CA, editors. Proceedings of Soil
Dynamics and Earthquake Engineering VII. Southampton, Boston,
USA: Computational Mechanics Publications, 1995. p. 55766.
[16] Watts GR. Vehicle generated ground-borne vibration alongside speed
control cushions and road humps. In: Proceedings of the Conference
on Ground Dynamics and Man-made Processes. Institution of Civil
Engineers, London, November 1997.
[17] Clemente P, Rinaldis D. Protection of a monumental building against
trafc-induced vibrations. Soil Dynamics and Earthquake Engineering 1998;17:28996.
[18] Hanazato T, Ugai K, Mori M, Sakaguchi R. Three-dimensional analysis of trafc-induced ground vibrations. Journal of Geotechnical Engineering, ASCE 1991;117(8):30016.
[19] Krylov VV. Effects of layered ground on ground vibrations generated
by high-speed trains. In: Proceedings of the Conference on Ground
Dynamics and Man-made Processes. Institution of Civil Engineers,
London, November 1997.
[20] Madshus C, Kaynia AM, Harvik L, Holme, JK. A numerical ground
model for railway-induced vibration. In: Proceedings of the Conference on Ground Dynamics and Man-made Processes. Institution of
Civil Engineers, London, November 1997.

[21] Wiss JE. Construction vibrations: state-of-the-art. Journal of the


Geotechnical Engineering Division, ASCE 1981;107(2):16781.
[22] Crabb GI, Hiller DM, Wilson PE. Measurement and prediction of
groundborn vibration due to construction operations. In: Proceedings
of the Conference on Ground Dynamics and Man-made Processes.
Institution of Civil Engineers, London, November 1997.
[23] Longinow A. Monitoring construction vibrations. In: Dusenberry Do,
Davie JR, editors. Effects of construction on structures, ASCE.
Geotechnical Special Publication No. 84, 1998:6779.
[24] Woods RD. Dynamic effects of pile installations on adjacent structures, NCHRP 253. Washington, D.C.: National Academy Press, 1997
[86 pp., Transportation Research Board].
[25] Clough GW, Chameau J-L. Measured effects of vibratory sheetpile
driving. Journal of Geotechnical Engineering Division, ASCE
1980;106(10):108199.
[26] Linehan PW, Longinow A, Dowding CH. Pipe response to pile driving and adjacent excavation. Journal of Geotechnical Engineering,
ASCE 1992;118(2):30016.
[27] Kim D-S, Lee J-S. Source and attenuation characteristics of various
ground vibrations. In: Dakoulas P, Yegian M, Holtz B, editors.
Geotechnical earthquake engineering and soil dynamics III. ASCE.
Geotechnical Special Publication No. 75, 1998;2:150717.
[28] Attwell PB, Farmer IW. Attenuation of ground vibrations from pile
driving. Ground Engineering 1973;63(7):269.
[29] Moore PJ, Styles JR, Ho W-H. Vibrations caused by pile driving. In:
Prakash S, editor. Proceedings of the Third International Conference
on Recent Advances in Geotechnical Earthquake Engineering and
Soil Dynamics, vol. III. St. Louis, MO, 1995. p. 73741.
[30] Attewell PB, Selby AR, O'Donnell L. Estimation of ground vibration
from driven piling based on statistical analyses of recorded data.
Geotechnical and Geological Engineering 1992;10:4159.
[31] Woods RD, Jedele LP. Energy-attenuation from construction vibrations. In: Gazetas G, Selig ET, editors. Vibration problems in geotechnical engineering. Special Publication of ASCE, 1985:22946.
[32] Yang XJ. Evaluation of man-made ground vibrations. In: Prakash, S,
editor. Proceedings of the Third International Conference on Recent
Advances in Geotechnical Earthquake Engineering and Soil
Dynamics, vol. III. St. Louis, MO, 1995. p. 13458.
[33] Ramshaw CL, Selby AR, Bettess P. Computation of the transmission
of waves from pile driving. In: Proceedings of the Conference on
Ground Dynamics and Man-made Processes. Institution of Civil
Engineers, London, November 1997.
[34] Ramshaw CL, Selby AR, Bettess P. Computed ground waves due to
piling. In: Dakoulas P, Yegian M, Holtz B, editors. Geotechnical
earthquake engineering and soil dynamics III, ASCE. Geotechnical
Special Publication No. 75, 1998;2:14841495.
[35] Svinkin HP. Overcoming soil uncertainty in prediction of construction and industrial vibrations. In: Shackelford CD, Nelson PP, Roth
MJS, editors. Uncertainty in the geologic environment: from theory
to practice, ASCE. Geotechnical Special Publication No. 58, 1996;2:
117894.
[36] Skipp BO. Dynamic ground movements: man-made vibrations. In:
Attwell PB, Taylor RK, editors. Ground movements and their effects
on structures. Surrey: Surrey University Press, 1984. p. 381434
Chapter 13.
[37] Siskind D, Stagg MS, Kopp JW, Dowding CH. Structure response and
damage produced by ground vibration from surface mine blasting.
Bureau of Mines Report RI 8507, Twin Cities, MN, NTIS #PB81157000, 1980. 74pp.
[38] Deutches Institut fur Normung, DIN 4150: Parts 1, 2 and 3. Vibrations
in buildings: effects on structures. Provisional Standards Revised
Draft, Part 3, Berlin, 1983.
[39] British Standards Institution, BS 5228: Part 4. Noise control on
construction and open sites, Part 4. Code of practice for noise and
vibration control applicable to piling operations. BSI, London, 1992.
[40] British Standards Institution, BS 7385: Part 1. ISO 4866: 1990.
Evaluation and measurement for vibration in buildings. Part 1.

G.A. Athanasopoulos, P.C. Pelekis / Soil Dynamics and Earthquake Engineering 19 (2000) 371387

[41]

[42]

[43]
[44]

[45]

[46]

[47]

[48]

Guide for measurement of vibrations and evaluation of their effects on


buildings. BSI, London, 1990.
Studer J, Suesstrunk A. Swiss standard for vibrational damage to
buildings. In: Proceedings of the 10th International Conference on
Soil Mechanics and Foundation Engineering, vol. 3. Stockholm,
A.A., Balkema, 1981. p. 30712.
Massarsch KR, Broms BB. Damage criteria for small amplitude
ground vibrations. In: Prakash S, editor. Proceedings of the Second
International Conference on Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics, vol. II. St. Louis, MO, 1991. p.
14519.
Lacy HS, Gould JP. Settlements from pile driving in sands. In: Gazetas G, Selig ET, editors. Vibration problems in geotechnical engineering. Special Publication of ASCE, 1985:15273.
Dowding CH. Vibration induced settlement from blast densication
and pile driving. In: Yeung AT, Felio GY, editors. Vertical and horizontal deformations of foundations and embankments, ASCE, vol. 1.
Geotechnical Special Publication No. 40, 1994:80617.
Leznicki JK, Gaibrois RG, Esrig MJ. Displacement of landmark
building resulting from adjacent construction activities. In: Yeung
AT, Felio GY, editors. Vertical and horizontal deformations of foundations and embankments, ASCE, vol. 1. Geotechnical Special Publication No. 40, 1994:22232.
Leathers FD. Deformations in sand layer during pile driving. In:
Yeung AT, Felio GY, editors. Vertical and horizontal deformations
of foundations and embankments, ASCE, vol. 1. Geotechnical Special
Publication No. 40, 1994:25768.
Massarsch KR. Static and dynamic soil displacements caused by pile
driving. In: Proceedings of the Fourth International Conference on the
Application of Stress Wave Theory to Piles. The Hague, The Netherlands, 1992. p. 7784.
Kim DS, Drabkin S, Rokhvarger A, Laefer D. Prediction of low level

[49]

[50]
[51]

[52]
[53]
[54]
[55]

[56]

387

vibrations induced settlement. In: Yeung AT, Felio GY, editors.


Vertical and horizontal deformations of foundations and embankments, ASCE, vol. 1. Geotechnical Special Publication No. 40,
1994:80617.
Kim DS, Drabkin S. Factors affecting vibration induced settlement.
In: Proceedings of the Third International Conference on Recent
Advances in Geotechnical Earthquake Engineering and Soil
Dynamics, vol. III. St. Louis, MO, 1995:11115.
Drabkin S, Lacy H, Kim DS. Estimating settlement of sand caused by
construction vibration. Journal of Geotechnical Engineering, ASCE
1996;122(11):9208.
Drabkin S, Lacy H. Prediction of settlements of structures due to pile
driving. In: Dakoulas P, Yegian M, Holtz B, editors. Geotechnical
earthquake engineering and soil dynamics III, ASCE. Geotechnical
Special Publication No. 75, 1998;2:1496506.
Gierke HE, Goldman DE. Effects of shock and vibration on man. In:
Harris CM, editor. Shock and vibration handbook, 3rd ed. New York:
McGraw-Hill, 1988 [chap. 44].
British Standards Institution, BS 6472: 1992. Evaluation of human
exposure to vibration in buildings (1 Hz to 80 Hz). BSI, London,
1992.
American National Standards Institute, ANSI S3 29-1993. Guide to
the evaluation of human exposure to vibrations in buildings. Acoustical Society of America, New York, NY, 1993.
Kelley PL, Dellorusso SJ, Russo CJ. Building response to adjacent
excavation and construction. In: Dusenberry DO, Davie JR, editors.
Effects of construction on structures, ASCE. Geotechnical Special
Publication No 84, 1998:8097.
Oriard LL, Richardson TL, Akins KP. Observed high-rise building
response to construction blast vibrations. In: Gazetas G, Selig ET,
editors. Vibration problems in geotechnical engineering. Special
Publication of ASCE, 1985:20328.

Potrebbero piacerti anche