You are on page 1of 14

Allowable Distance from Impact Pile Driving to Prevent

Structural Damage Considering Limits in Different Standards


Amir Hamidi, A.M.ASCE1; Abtin Farshi Homayoun Rooz, M.ASCE2; and Majid Pourjenabi, S.M.ASCE3

Abstract: Pile driving has usually been the most prevalent method of pile installation, although the ground vibrations generated by pile driving
operations have a high damage potential for structures in the vicinity and are a major concern. Hence, the intention of this study is first to facili-
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

tate problems associated with the ground vibration issues and then to evaluate the allowable distance of different structures adjacent to an impact
pile driving site to prevent structural damage based on the acceptable vibration levels of reliable standards and a report in terms of peak particle
velocity (PPV). In this regard, the numerical simulation of a continuous impact pile driving process for a case study is executed from the ground
surface using the ALE (arbitrary Lagrangian-Eulerian) adaptive mesh method rather than common discontinuous pile driving models. Based on
the results, peak vertical velocity is the most appropriate approach for the determination of PPV and the investigation of the ground vibrations
induced during construction activities. Furthermore, the most comprehensive standard is concluded to assess transient vibrations caused by con-
struction activities. For practical applications, maximum peak allowable PPV using the standard resultant was presented to prevent structural
damage to different structures. DOI: 10.1061/(ASCE)SC.1943-5576.0000354. © 2017 American Society of Civil Engineers.
Author keywords: Pile driving; Ground vibrations; Structural damage; Peak particle velocity (PPV); Numerical simulation; Maximum
peak allowable PPV.

Introduction the soil profile near the pile as well as the proximity of adjacent sur-
face and buried structures is essential to avoid structural damage.
Through the ages, pile driving has been the most efficient method for Moreover, Athanasopoulos and Pelekis (2000), Thandavamoorthy
pile installation by applying consecutive impacts of a hammer on a (2004), Svinkin (2006), Madheswaran et al. (2009), Zhang et al.
precast pile head. The rapid growth of population along with con- (2013), and Grizi et al. (2016) acknowledged that the ground vibra-
struction have led to the proximity of various structures to each other, tions originating from pile driving are quite severe and have a signifi-
especially in urban areas; therefore, concerns over possible irrepara- cant damage potential for the nearby structures. Hence, the acceptable
ble damage to structures in the vicinity of pile driving has been dra- vibration limits in standards for various structures are perused
matically increased. Hence, a number of standards have proceeded to through several studies, such as Hiller and Hope (1998), White et al.
regulate criteria for avoiding damage to the nearby structures by (2002), Jones and Stokes Associates (2004), Sylvestre-Williams
increasing the knowledge and understanding of construction vibra- (2011), Ekanayake et al. (2013), and Massarsch and Fellenius (2015).
tions [Swiss Standard SN640312 (SNV1992); Eurocode 3 (CEN It is virtually impossible to completely perceive the ground
1993); Swedish Standard SS 02 52 11 (SIS 1999); German Standard vibration problems unless the entire chain of vibration transmission
DIN 4150-3 (DIN 1999)] The prediction of construction vibrations is considered; hence, in Fig. 1 the ground vibration transmission
prior to inception has been necessitated by these standards. process is illustrated during pile driving operations. In the process,
Retrospectively, Wiss (1981) surveyed the detrimental effects of hammer impact is applied to the pile head, and then the vibrations
construction vibrations by considering the sources of construction are propagated at the pile-soil interface into the surrounding soil
vibrations, transmitting medium, and effects on the vibration through pile-soil interaction. Wave propagation occurs throughout
receivers. Uromeihy (1990) proceeded with the ground vibration the soil layers in the shape of transient vibration from impact pile
measurements with a particular reference to pile driving; thus, the driving, or continuous vibration from vibratory pile driving.
effect of ground vibrations on buildings was investigated in large- Finally, the ground vibrations impact any structure in their path, and
scale tests. Woods (1997) investigated the dynamic effects of pile in- structural damage can occur through soil-structure interaction.
stallation on the adjacent structures and concluded that the pile driv- Thus far, numerical modeling of pile driving has been mostly ex-
ing vibrations can be a problem with all kinds of pile drivers and all ecuted through discontinuous pile driving. In other words, a mod-
kinds of driven piles. It was recommended that detailed knowledge of eled pile was embedded in a prebored hole in the soil (desired
depth) and then just a few hammer impacts or a small displacement
1
Professor, School of Engineering, Kharazmi Univ., P.O. Box 15614, was applied; that is, continuous pile driving from the ground surface
Tehran, Iran (corresponding author). E-mail: hamidi@khu.ac.ir has not been performed (Mabsout and Tassoulas 1994; Ramshaw
2
Senior Geotechnical Engineer, School of Engineering, Kharazmi 2002; Masoumi et al. 2009; Henke 2010; Dijkstra et al. 2011;
Univ., P.O. Box 15614, Tehran, Iran. E-mail: std_abtinfhr@khu.ac.ir Zhang et al. 2013; Lupiezowiec et al. 2014; Liyanapathirana and
3
Ph.D. Candidate, School of Engineering, Kharazmi Univ., P.O. Box Ekanayake 2016). Although the discontinuous pile driving leads to
15614, Tehran, Iran. E-mail: std_pourjenabi@khu.ac.ir
a considerable reduction of analysis time, the soil conditions and
Note. This manuscript was submitted on April 24, 2017; approved on
August 24, 2017; published online on November 28, 2017. Discussion pe- the results were relatively far from field measurements because of
riod open until April 28, 2018; separate discussions must be submitted for not simulating the stress and strain changes of surrounding soil dur-
individual papers. This paper is part of the Practice Periodical on ing pile driving from the ground surface. Therefore, the most supe-
Structural Design and Construction, © ASCE, ISSN 1084-0680. rior part of the numerical modeling in this paper is to simulate the

© ASCE 04017029-1 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


continuous impact pile driving process from the ground surface to and difficulties that have particular importance in the beginning
the final pile installation depth just identical to real works rather steps. Hence, the understanding of the following concepts seems
than the prevalent discontinuous pile driving. Subsequently, deter- certainly worthwhile and should be observed.
mination of the allowable distance of different structures in a case First and foremost, it is essential to find out which parameter is
study of the impact pile driving site was performed based on the the best option with which to investigate the ground vibrations
maximum acceptable level of various authentic standards and a induced by various sources. Thus far, the ground vibrations have
report. Moreover, to overcome the common beginning difficulties been evaluated mostly in terms of peak particle displacement (PPD),
of studying ground vibration issues, basic crucial factors, including peak particle velocity (PPV), and peak particle acceleration (PPA)
the best parameter of evaluating the ground vibrations, the vibration by different researchers (Uromeihy 1990; Mabsout and Tassoulas
construction types, and the best approach of maximum vibration 1994; Athanasopoulos and Pelekis 2000; Thandavamoorthy 2004;
determination, are meticulously investigated in advance. Finally, a Çelebi et al. 2009; Ekanayake et al. 2013; Svinkin 2015; Pourjenabi
comprehensive guideline considering the limits of standards was and Hamidi 2015; Rezaei et al. 2016; Grizi et al. 2016), which are
presented to avoid structural damage in practical applications. the maximum displacement, velocity, and acceleration that a soil
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

particle experiences during a vibration event, respectively. A num-


ber of researchers have shown better consistency of PPV to the ini-
Basic Crucial Factors tiation of damage in the structures compared with PPD and PPA,
and PPV has been introduced as the most appropriate descriptor for
To investigate the ground vibrations generated by diverse sources, evaluation of the ground vibrations (New 1986; Kennedy 1990;
researchers are indeed confronted with some common ambiguities Massarsch and Broms 1991; Jones and Stokes Associates 2004;

Fig. 1. Vibration transmission process during pile driving

15
Impact Pile Driving (Transient Vibrations)
Vibration Velocity: mm/sec

10 Vibratory Pile Driving (Continuous Vibrations)

-5

-10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time: sec

Fig. 2. Comparison of vibrations for a soil particle adjacent to impact and vibratory pile driving

© ASCE 04017029-2 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


Svinkin 2004; Abdel-Rasoul and Mohamed 2006). It is worth noting sources: natural phenomena (e.g., earthquake, landslide, volcanic
that the standards mostly address the ground vibrations based on eruptions, tsunami, meteorite impact, etc.) and man-made activities
PPV that corroborate the conclusion. (mostly construction activities). It is worth noting that the ground
Second, it is also crucial to figure out the type of ground vibra- vibrations generated from construction activities can be classified
tions. Generally, ground vibrations can originate from two main into transient and continuous vibrations.
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Three orthogonal directions (X, Y, and Z) of vibration measurements as velocity components of a soil particle (VR, VT, and VV)

Table 1. Maximum Allowable PPV Values for Transient Vibrations to Prevent Structural Damage in Authentic Standards and Reports

Reference Frequency range (Hz) Structure and condition Maximum PPV [mm/s (in./s)]
AASHTO (1990) — Engineered structures, without plaster 25.4–38.1 (1–1.5)
Residential building in good repair with gypsum 10.16–12.7 (0.4–0.5)
board walls
Residential buildings, plastered walls 5.08–7.62 (0.2–0.3)
Historic sites or other critical locations 2.54 (0.1)
Swiss Standard 10–30 (I) Buildings in steel or reinforced concrete, such as 30.48 (1.2)
SN640312 (SNV 1992) factories, retaining walls, bridges, steel towers, open
channels, underground chambers, and tunnels with
and without concrete alignment
(II) Buildings with foundation walls and floors in 17.78 (0.7)
concrete, walls in concrete or masonry, stone ma-
sonry retaining walls, underground chambers and
tunnels with masonry alignment, conduit in loose
material
(III) Buildings as mentioned previously but with 12.7 (0.5)
wooden ceilings and walls in masonry
(IV) Construction very sensitive to vibration, objects 7.62 (0.3)
of historic interest
British Standard 4–15 Unreinforced or light framed structures, residential 15–20 (0.6–0.8)
or light commercial type buildings
BS 7385-2 (BSI 1993) >15 Unreinforced or light framed structures, residential 20–50 (0.8–2)
or light commercial type buildings
All frequencies Reinforced or framed structures–industrial and 50 (2)
heavy commercial buildings
Eurocode 3 (CEN 1993) — Buried services 40 (1.6)
Heavy industrial 30 (1.2)
Light commercial 20 (0.8)
Residential 10 (0.4)
Ruins, building of architectural merit 4 (0.15)
FHWA Dynamic Compaction — Buried pipes and mains 76 (3)
(Lukas 1995)
German Standard 0–10 Commercial–industrial 20 (0.8)
DIN 4150-3 (DIN 1999) 10–50 20–40 (0.8–1.6)
50–100 40–50 (1.6–2)
0–10 Residential 5 (0.2)
10–50 5–15 (0.2–0.6)
50–100 15–20 (0.6–0.8)
0–10 Sensitive–historic 3 (0.1)
10–50 3–8 (0.1–0.3)
50–100 8–10 (0.3–0.4)

© ASCE 04017029-3 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


The transient (single-event) vibrations rapidly reach a peak independent of their occurrence time. It is evident that the simul-
value and then are subjected to a damping effect until it disap- taneous occurrence of maximum velocity components for a soil
pears. The transient vibration sources may include only a single particle during construction work is far from reality. Therefore,
event, such as an explosion, or numerous single events, such as PPV determination using Eq. (2) would be definitely conserva-
impact pile driving, dynamic compaction, and so forth. The con- tive; nonetheless, several researchers have applied the approach
tinuous (steady-state) vibrations generate uninterrupted oscilla- to their investigations (Head and Jardine 1992; Hiller and Hope
tion for a defined period with manifold cycles. The common con- 1998; Rockhill et al. 2003).
tinuous vibration sources are vibratory pile driving, roadway and 3. Peak unidirectional velocity: In this approach, PPV is equal to
railway traffic, underground railway, heavy construction equipment the component with the highest recorded value during the entire
(e.g., jackhammer, scraper, dump truck, bulldozer, etc.), excavation construction time. Actually, PPV can be each of the three ve-
(e.g., tunnel boring machinery), demolition activity, static and vibra- locity components independent of the two other components.
tory compaction equipment (e.g., vibratory rollers, wheeled and
sheep foot rollers), factory machinery, and so forth. PPV ¼ Max ðVRmax ; VVmax ; VTmax Þ (3)
Among the construction sources, pile driving is the most prevalent
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

source of construction vibrations (Karlsson 2013). It is also important Several valid sources have used the approach [BS 7385-2:1993
to note that the allowable vibration limit in standards was determined (BSI 1993), DIN 4150-3 (DIN 1999), and Zhang et al. (2013)].
based on the generated vibration type (transient/continuous); hence, 4. Peak vertical velocity: The maximum vertical component of
it can be concluded that the nature of these two vibration types is velocity obtained from measurements or computations is con-
completely different. Thus, to make the difference clear, in Fig. 2 sidered as PPV according to Eq. (4)
impact pile driving transient vibrations and vibratory pile driving
continuous vibrations are compared. As illustrated in Fig. 2, tran- PPV ¼ VVmax (4)
sient vibrations often have a higher intensity than continuous vibra-
tions. However, even though standards have determined higher lim- In contrast to the former approaches, this approach has been
its for transient vibrations than for continuous vibrations, continuous widely applied by a number of researchers as well as standards
vibrations have greater potential to cause structural damage. Both [Attewell and Farmer (1973); Martin (1980); Massarsch and
vibration types, especially those induced by pile driving, have great Broms (1991); Swiss Standard SN640312 (SNV 1992);
potential to cause structural damage and to annoy residents in adja- Swedish Standard SS 02 52 11 (SIS 1999); Athanasopoulos
cent structures. and Pelekis (2000); Thandavamoorthy (2004); Ahlquist and
The type of ground vibration originating from impact pile driving Enggren (2006); Whenham et al. (2009); Serdaroglu (2010);
is called a transient vibration. To determine the maximum PPV dur- Schumann and Grabe (2011); Athanasopoulos et al. (2013);
ing construction work at a site through measurements or in numeri- Massarsch and Fellenius (2015); Grizi et al. (2016)].
cal analyses, researchers always face several approaches that make Of the four represented approaches, the last approach (peak vertical
the determination extremely ambiguous. Therefore, great effort was velocity) is the best option because
expended in the current study to gather and evaluate all the existing 1. It was used in most of the previous research studies,
approaches and then, more importantly, identify the valid resources 2. There is better correlation with damage inception,
of using each approach to decide which one is the most appropriate
option. To assess the ground vibrations during a construction pro-
ject, such as pile driving, first the measurements and computations
Table 2. Maximum Allowable PPV Values for Transient Vibrations to
of the vibration velocity at different distances from the source are Prevent Structural Damage in the Swedish Standard SS 02 52 11 (SIS 1999)
recorded simultaneously in three orthogonal directions of X, Y, and
Z as the velocity components (VR, VV, and VT) at any moment, as Factor Condition Value
shown in Fig. 3. Indeed, VR, VV, and VT are known as radial (or longi- V0 [mm/s (in./s)] Bedrock 15 (0.6)
tudinal), vertical, and tangential (or transverse) velocity compo- Glacial till 12 (0.5)
nents, in which VR must be toward the source. The approaches for Clay, silt, sand, or gravel 9 (0.35)
PPV determination in the scientific studies are as follows: Fb Heavy structures, such as bridges, quay 1.7
1. Peak vector sum (PVS): The most well-known approach for walls, defense structures, etc.
PPV determination in scientific literature is the maximum re- Industrial or office buildings 1.2
sultant of velocity components that a soil particle experiences Normal residential buildings 1
at the same time during construction work, as shown in Eq. (1) Especially susceptible buildings and 0.65
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi buildings with high value or structural
PPV ¼ Max ðVR Þ2 þ ðVV Þ2 þ ðVT Þ2 (1) elements with spans, e.g., church or
museum
Despite being the popular definition, only a few researchers Historic buildings in a sensitive state as 0.5
have used this approach (Uromeihy 1990; Ho and Tan 2003; well as certain sensitive ruins
Liden 2012; Ekanayake et al. 2013). Fm Reinforced concrete, steel, or wood 1.2
2. Square root of squares sum (SRSS): Another approach, as can Unreinforced concrete, bricks, concrete 1
be seen in Eq. (2), is the resultant of maximum velocity compo- blocks with voids, light-weight concrete
nents aside from considering the occurrence time. elements, masonry
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Light concrete blocks and plaster 0.75
PPV ¼ ðVRmax Þ2 þ ðVVmax Þ2 þ ðVTmax Þ2 (2) Limestone 0.65
Fg Buildings founded on toe-bearing piles 1
Buildings founded on shaft-bearing piles 0.8
where VRmax , VVmax , and VTmax = maximum recorded values of
Spread footings, raft foundations 0.6
velocity components during the entire construction period

© ASCE 04017029-4 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


3. It is relatively easier to measure or compute, and transient vibrations was extracted from well-known and authentic
4. There are close results to the component resultant (PVS). standards including AASHTO (1990), Swiss Standard SN640312
This approach has been used in the numerical analysis to (SNV 1992), British Standard BS 7385-2 (BSI 1993), Eurocode 3
follow. (CEN 1993), and German Standard DIN 4150-3 (DIN 1999) as
well as the report Federal Highway Administration (FHWA)
Dynamic Compaction (Lukas 1995), as shown in Table 1. Also,
Maximum Allowable PPV in Standards and Reports Swedish Standard SS 02 52 11 (SNV 1999) as one of the available
comprehensive standards has considered maximum allowable
Because of the abundant advantages of pile installation by pile PPV (millimeters per second) depending on some parameters, i.
driving compared with other methods, widespread use of the pile e., vibration velocity in different soil types (V0), building type
driving method for pile installation is understandable. Despite the (Fb), building materials (Fm), and foundation type (Fg), which is
notable advantages of the pile driving method, the adverse effect computable using Eq. (5) and Table 2
of ground vibrations on the surrounding structures is the only rea-
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

son for limiting the method. Therefore, many researchers and PPV ¼ V 0  Fb  Fm  Fg (5)
standards have proceeded with the determination of a maximum
acceptable PPV value to prevent the possible structural damage.
They have unanimously expressed the evaluation of anticipated
ground vibrations to be less than the maximum permissible deter- Numerical Modeling and Verification
mined PPV value before initiation of any construction operation
is absolutely indispensable. Indeed, the maximum determined To conduct the numerical simulation of the continuous impact
PPV values correspond to the damage threshold of the structures. pile driving process, the finite-element software Abaqus 6–14-3
Hence, in the present research, the maximum allowable PPV for using an explicit dynamic integration scheme was utilized, in

Fig. 4. Geometry, boundary condition, and finite-element mesh of the axisymmetric model

Table 3. Pile and Soil Properties Used in the Numerical Modeling (Data from Wiss 1981, © ASCE)

Diameter Cohesion Friction angle Density [kg/m3 Elastic modulus Poisson’s


Type Material [m (ft)] Length [m (ft)] [kPa (psi)] (degrees) (lb/ft3)] [MPa (ksi)] ratio
Pile Concrete 0.5 (1.64) 10 (32.8) — — 2,500 (156) 40,000 (5,801) 0.25
Soil Sandy clay — — 15 (2.18) 25 2,000 (125) 80 (11.6) 0.4

© ASCE 04017029-5 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


which an axisymmetric model space was considered with the pile soil are imminent during the pile penetration process. As a result, the
centerline located along the axis of the symmetry. The ongoing nodes have an ability to move independently, and it is possible to
numerical model was verified by a reported case study from Wiss make the soil elements smaller during pile penetration if needed with-
(1981). out changing the geometry.
Fig. 4 illustrates the geometry of pile and soil models. The Indeed, numerical modeling of the pile penetration process
soil is homogeneous sandy clay, and the concrete pile is closed from the ground surface to a desired depth using the classical
end with a 0.5 m (1.64-ft) diameter, 10 m (32.8-ft) length, and FEM is infeasible due to the expected large deformations of the
conical tip angle equal to 60° in which the properties are fully soil around the penetrating pile, which leads to excessive element
represented in Table 3. The elastic and Mohr-Coulomb failure distortions and cessation of the analysis afterward. Therefore, to
criterion were considered for the pile and soil behaviors, respec- cope with the difficulties, the ALE adaptive mesh method was
tively. As shown in Fig. 4, the soil damping was considered 7% and applied to the model.
the artificial boundary of five subzones with 7.5–17.5% Rayleigh Because the pile driver hammer was a BSP 357 hydraulic impact
damping was applied to the right side boundary of the model to pre- hammer, for each hammer impact the hammer force-time curve of
vent wave reflection. For the bottom boundary, the soil depth was Deeks and Randolph (1993) was applied to the pile head. Because
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

extended to the size that waves can completely damp through its natu- of the designation of an elastic behavior for the pile, the concen-
ral damping before reaching the bottom boundary. The finite-element trated hammer force was converted to normal stress in advance, as
mesh of the model is also demonstrated in Fig. 4. Four-node quadratic delineated in Fig. 5. Meanwhile, a 1-s time interval was considered
structured elements (CAX4Rs) were used to mesh the soil and pile for each hammer impact.
models. Moreover, as shown in Fig. 4, ALE (arbitrary Lagrangian- The pile-soil interaction was defined through a surface-to-
Eulerian) adaptive mesh was applied to the limited soil zone meas- surface contact of master-slave technique, and its properties were
uring 3  13 m (9.84  42.65 ft) in which the large displacements of allocated through normal behavior with hard contact ability and tan-
gential behavior using a penalty friction formulation. Also, the fric-
tion coefficient was reasonably considered equal to 0.35 based on
the materials of the pile and soil.
To install the pile in the soil through continuous impact pile driv-
ing in the ongoing numerical analyses, at first, a short time interval
was used on the pile for initial static penetration into the soil due to
its weight right after application of the earth gravity acceleration to
create in situ stresses of the pile and soil. Then, consecutive hammer
impacts on the pile head were performed based on the previously
mentioned hammer stress history for reaching the pile to a 10 m
(32.8-ft) final penetration depth.
Finally, to verify the numerical model of continuous impact pile
driving from the ground surface to the final penetration depth of
10 m (32.8 ft), the field measurements reported by Wiss (1981)
were used to validate the model in terms of PPV at the different hor-
izontal distances of the ground surface, as shown in Fig. 6.
According to Fig. 6, the numerical results are in full accord with the
Fig. 5. Impact force of the BSP 357 pile driver hammer in terms of measured data. To ensure the numerical results, the comparison of
normal stress calculated and measured PPV values was surveyed in a wide dis-
tance range of 1–45 m (3.28–147.64 ft).

1000
Measured by Wiss (1981)
Peak Particle Velocity (PPV): mm/s

Current Study

100

10

1
1 2 4 8 16 32 64
Horizontal Distance (r): m

Fig. 6. Comparison of computed and measured PPV values induced by impact pile driving for soil particles at the ground surface and located at dif-
ferent distances from the pile centerline

© ASCE 04017029-6 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


Comparison of Vibration Components with the PVS with the PVS of the components. For measurement of the maxi-
mum vibration velocity during the impact pile driving process
As previously mentioned, vibration velocity components of a from the ground surface to the final penetration depth of 10 m
soil particle adjacent to a pile driving site are vertical, radial, (32.8 ft), which needed 353 hammer impacts, 1,000 records
and transverse, in which vertical and radial components have were concluded to perform after each hammer impact during the
always been considered as the significant components in previ- 1-s time interval before application of the next hammer impact,
ous studies, and transverse vibration has usually been ignored namely, a velocity record in the 0.001-s time interval. Regarding
because it is very small. Hence, in almost all cases it is sufficient the figure, it can be stated that peak vertical velocity (the fourth
to measure only vertical and radial vibration and ignore the prevoiusly mentioned approach) is in close accordance with PVS
transverse component of the vibration (Uromeihy 1990; (first approach), but peak radial velocity is 40–70% smaller than
Athanasopoulos and Pelekis 2000; Hanson et al. 2006). peak vertical velocity and PVS. Consequently, the peak vertical
Therefore, in Fig. 7 peak vertical and radial velocity compo- velocity can be concluded as the best descriptor to investigate the
nents of the continuous pile driving case study are compared ground vibrations.
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

1000
Peak Vector Sum
Peak Particle Velocity (PPV): mm/s

Peak Vertical Velocity

Peak Radial Velocity


100

10

1
1 2 4 8 16 32 64
Horizontal Distance (r): m

Fig. 7. Comparison between two vibration components of vertical (peak vertical velocity) and radial (peak radial velocity) with PVS

1000
Peak Particle Velocity (PPV): mm/s

100

(I)
(II)
(III)
10 (IV)

1
1 2 4 8 16 32 64
Horizontal Distance (r): m

Fig. 8. Allowable level of ground vibrations for different structures adjacent to the impact pile driving to prevent structural damage based on the
Swiss Standard SN640312 (SNV 1992)

© ASCE 04017029-7 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


Allowable Distance of Structures for the Considered structures from the pile driving site can be determined; otherwise,
Case Study the generated vibrations should be reduced to the maximum allow-
able PPV levels through an efficient method to avoid structural
To avoid likely damage of different structures adjacent to a vibra- damage. The prevalent efficient methods are active or passive wave
tion generator source, standards have determined the threshold barriers (open or in-filled trenches), preboring, and water jetting. In
guidelines for ground vibrations in terms of PPV. In other words, the following subsections, the maximum allowable distance of
based on the regulated levels, a maximum safe distance of various structures from the pile driving case study of Wiss (1981) is investi-
gated based on the limits of authentic standards and reports. Also,
Table 4. Maximum Allowable Distance of Various Structures from the the limits in the following investigations are just for transient sour-
Pile Driving Site to Prevent Structural Damage Based on the Swiss ces, such as impact pile driving.
Standard SN640312 (SNV 1992)

Maximum Maximum
Swiss Standard SN640312 (SNV 1992)
allowable PPV allowable To prevent probable damage of structures adjacent to a pile driving
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

Number Structural conditions [mm/s (in./s)] distance [m (ft)] site, Swiss Standard SN640312 (SNV 1992) determined the maxi-
I Buildings in steel or rein- 30.48 (1.2) 7 (23) mum allowable PPV values based on different conditions of differ-
forced concrete, such as ent structures, as shown in Table 1. Based on this standard, more re-
factories, retaining walls, sistant structures less susceptible to vibrations can withstand a
bridges, steel towers, open higher level of vibration without damage; thus, the maximum
channels, underground allowable distance from the source will be smaller (closer to the
chambers and tunnels with vibration source). To reveal the maximum allowable distance of
and without concrete structures from the pile driving site in the case study based on Swiss
alignment Standard SN640312 (SNV 1992), the ground vibrations obtained
II Buildings with foundation 17.78 (0.7) 10.7 (35.1) through continuous impact pile driving simulations as well as maxi-
walls and floors in concrete, mum PPV levels in the standard are delineated in Fig. 8. Evidently,
walls in concrete or masonry, the intercept of the ground vibrations diagram and the maximum
stone masonry retaining allowable PPV line is the maximum allowable distance from the
walls, underground chambers pile driving site, as presented in Table 4. With respect to the results
and tunnels with masonry of maximum allowable distances based on the Swiss Standard
alignment, conduit in loose SN640312 (SNV 1992), it can be concluded that weaker structures
material and structures susceptible to vibrations (IV) require being about
III Buildings as mentioned 12.7 (0.5) 13.5 (44.3) two to three times farther away from the pile driving site than the
previously but with wooden more resistant structures (I and II).
ceilings and walls in
masonry Eurocode 3 (CEN 1993)
IV Construction very sensitive 7.62 (0.3) 19.2 (63)
to vibration, objects of Eurocode 3 (CEN 1993) has undoubtedly been the most prevalent
historic interest and well-known standard in the ground vibrations assessment. It
has provided limits in terms of PPV for various structures exposed

1000
Peak Particle Velocity (PPV): mm/s

100

Buried Services
Heavy Industrial
Light commercial

Residenal
10

Ruins, buildings of architectural merit

1
1 2 4 8 16 32 64
Horizontal Distance (r): m

Fig. 9. Allowable level of ground vibrations for different structures in the vicinity of the impact pile driving to prevent structural damage based on
Eurocode 3 (CEN 1993)

© ASCE 04017029-8 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


to transient or continuous ground vibrations. In other words, struc- most of the transmitted energy by Rayleigh waves, the maximum
tural damage thresholds were defined by the standard; thus, the limi- PPV occurred at the ground surface. Regarding the maximum
tation range was from a maximum allowable PPV of 4 mm/s (0.15 allowable PPV of 40 mm/s (1.6 in./s) for buried services, the
in./s) for structures susceptible to ground vibrations, such as historic maximum allowable distance was obtained as 5.6 m. However,
structures, to 40 mm/s (1.6 in./s) for buried services, such as water in practice, buried services are usually located in different
pipes, gas or oil pipelines, electricity, and telecommunication ground depths, so their PPV value is less than the ground sur-
cables. The maximum allowable PPV values for transient vibrations face. Therefore, the maximum allowable distance for buried
were assigned as twice greater than the continuous vibrations in the services seems conservative.
standard. As shown in Fig. 9, by overlaying the criteria of Eurocode
3 (CEN 1993) for transient vibrations on the resultant ground vibra-
FHWA Dynamic Compaction Report (Lukas 1995)
tions obtained from the continuous impact pile driving simulation
of the case study, the maximum allowable distance can be computed Ground vibrations due to the dynamic compaction operations are
(Table 5). According to the results of maximum allowable distances transient vibrations. Thus, it is feasible to apply the limit of the
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

from the pile driving site, the residential buildings should be a dis- authentic report FHWA Dynamic Compaction (Lukas 1995) to the
tance greater than 16 m (52.5 ft) from the pile driving site to be safe; ground vibrations induced by pile driving. The report declares that
otherwise, ground vibrations should be reduced to the maximum considering PPV measurements on the ground over buried utilities,
allowable PPV value of 10 mm/s (0.4 in./s) through an efficient 76 mm/s (3 in./s) is the safe level for pipes and mains. Hence, in
method. Moreover, the ruins and structures susceptible to ground Fig. 10 the maximum allowable safe level of PPV is superimposed
vibrations require being about two times farther from the pile driv- on the results of the propagated ground vibrations through pile driv-
ing site than residential buildings. Additionally, due to carrying ing of the case study, and the maximum allowable distance from the

Table 5. Maximum Allowable Distance of Various Structures from the Pile Driving Site to Prevent Structural Damage Based on Eurocode 3 (CEN 1993)

Number Structure type Maximum allowable PPV [mm/s (in./s)] Maximum allowable distance [m (ft)]
1 Buried services 40 (1.6) 5.6 (18.4)
2 Heavy industrial 30 (1.2) 7.1 (23.3)
3 Light commercial 20 (0.8) 9.7 (31.8)
4 Residential 10 (0.4) 16 (52.5)
5 Ruins, building of architectural merit 4 (0.15) 30.2 (99.1)

1000
Peak Particle Velocity (PPV): mm/s

100 Buried pipes and mains

10

1
1 2 4 8 16 32 64
Horizontal Distance (r): m

Fig. 10. Allowable level of ground vibrations for different structures adjacent to the impact pile driving to prevent structural damage based on the
FHWA Dynamic Compaction Report (Lukas 1995)

Table 6. Maximum Allowable Distance from the Pile Driving Site to Prevent Structural Damage Based on the FHWA Dynamic Compaction Report (Lukas
1995)

Number Structure type Maximum allowable PPV [mm/s (in./s)] Maximum allowable distance [m (ft)]
1 Buried pipes and mains 76 (3) 4 (13.1)

© ASCE 04017029-9 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


pile driving site is concluded at 4 m (13.1 ft), as shown in Table 6. It are actually located in the ground and depend on their embedment
is evidently important that the FHWA criteria for buried structures, depth, these structures certainly experience much fewer vibrations
similar to Eurocode 3 (CEN 1993), are based on the ground vibra- than the ground surface vibration level. For example, in the pile
tion measurements at the ground surface, whereas buried structures driving site considering a maximum allowable distance equal to
4 m (13.1 ft) for a gas pipe located at depths of 1, 10, and 20 m
(3.28, 32.8, and 65.6 ft) sounds irrational. Accordingly, considering
Table 7. Maximum Allowable Distance of Various Structures from the a maximum constant level for buried structures in various ground
Pile Driving Site to Prevent Structural Damage Based on the Swedish depths is definitely a conservative approach. The solution to the
Standard SS 02 52 11 (SIS 1999)
problem suggested in this paper is to use the standards’ application
Maximum Maximum of a reduction coefficient to the maximum allowable PPV based on
allowable PPV allowable the embedment depth of a buried structure.
Number Conditions [mm/s (in./s)] distance [m (ft)]
1 Sandy soil–bridge– steel– 14.7 (0.6) 12.5 (41) Swedish Standard SS 02 52 11 (SIS 1999)
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

founded on shaft–bearing Because it considers several factors affecting the ground vibrations,
piles e.g., soil type, structure type, structure materials, and foundation
2 Clayey soil–office building– 7.78 (0.3) 18.9 (62) type, Swedish Standard SS 02 52 11 (SIS 1999) is undoubtedly the
reinforced concrete–raft
most comprehensive standard. However, it is crucial to note that the
foundation
standard criteria are consequences of long-term practical experien-
3 Clayey soil–historic build- 4.5 (0.18) 28 (91.9)
ces with transient vibrations generated by pile driving, sheet piling,
ing–masonry materials–
and soil compaction in Sweden; thus, it should be carefully applied
founded on toe-bearing piles
in other areas. To put it in practical use, five different conditions are
4 Sandy soil–residential 3.6 (0.14) 31.7 (104)
considered in Table 7, and then the maximum allowable PPV is cal-
building–unreinforced
culated through the relevant relationship. Finally, the maximum
concrete–raft foundation
allowable distances can be readily concluded using superimposition
5 Sandy soil–church–light 2.63 (0.1) 36.9 (121)
of limits and the ground vibrations induced by the impact pile driv-
concrete blocks and plaster–
ing, as summarized in Table 7. According to the results, unlike most
raft foundation
of the standards, structure type is not the only important factor in

Table 8. Maximum Allowable Distance of Various Structures from the Pile Driving Site to Prevent Structural Damage Based on German Standard DIN
4150-3 (DIN 1999)

Number Building type Maximum allowable PPV [mm/s (in./s)]a Maximum allowable distance [m (ft)]
I Commercial–industrial 20 (0.8) 9.7 (31.8)
II Residential 5 (0.2) 25.5 (83.7)
III Sensitive–historic 3 (0.1) 34.6 (113.5)
a
For the range of vibration frequencies between 0–10 Hz.

1000
Peak Particle Velocity (PPV): mm/s

100

Commercial / Industrial

10
Residenal
Sensive / Historic

1
1 2 4 8 16 32 64
Horizontal Distance (r): m

Fig. 11. Allowable level of ground vibrations for different structures adjacent to the impact pile driving to prevent structural damage based on the
German Standard DIN 4150-3 (DIN 1999)

© ASCE 04017029-10 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


the determination of the maximum allowable PPV to prevent struc-

Maximum
7.62 (0.3)

20 (0.8)

50 (2)

76 (3)
tural damage. Consequently, the Swedish standard is currently
superior to the other standards due to using a number of significant
factors affecting the ground vibrations.

Minimum
2.54 (0.1)

5 (0.2)

20 (0.8)

40 (1.6)
German Standard DIN 4150-3 (DIN 1999)
To avoid structural damage from the ground vibrations of construc-
tion activities, the German standard was initially released in 1970

German Standard
and has since undergone a number of revisions. Finally, the latest

DIN 4150-3
(DIN 1999)
version was issued as German Standard DIN 4150-3 (DIN 1999),

3 (0.1)

5 (0.2)

20 (0.8)


which provides guideline levels for the ground vibrations. As
shown in Table 1, the German standard limited the ground vibra-
tions based on the building type and the frequency content of
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

vibrations as well. It is also important that, according to the stand-

FHWA Dynamic
ard, commercial and industrial buildings can withstand higher

(Lukas 1995)
Compaction
vibration levels than residential, sensitive, and historic buildings.

76 (3)
Furthermore, the vibration levels with a frequency range of 0–


10 Hz were adopted in this paper because the pile driving opera-
tions usually have a low vibration frequency. Eventually, as pre-
sented in Table 8, the maximum allowable distance of different

20–30 (0.8–1.2)
buildings from the impact pile driving site for the case study was

(CEN 1993)
Eurocode 3
Table 9. Maximum and Minimum Peak Allowable PPV of the Standard Resultant for Transient Vibrations, such as Impact Pile Driving

4 (0.15)
extracted through superimposition of the results of ground vibra-

10 (0.4)

40 (1.6)
tions induced by impact pile driving and the maximum allowable
PPV levels of the German standard, as illustrated in Fig. 11. Thus,
it can be concluded that residential, sensitive, and historic build-
ings should be three to four times farther from the pile driving site
British Standard

5.08–12.7 (0.2–0.5) 12.7–17.78 (0.5–0.7) 15–20 (0.6–0.8)


than commercial/industrial buildings.

(BSI 1993)
BS 7385-2

50 (2)


Allowable Vibration Level Considering the
Standard Resultant

25.4–38.1 (1–1.5) 17.78–30.48 (0.7–1.2)


The level of permissible PPV was thoroughly surveyed based on
Swiss Standard

7.62 (0.3)

each standard beforehand. Since each standard has addressed only a


(SNV 1992)
SN640312

few structures or conditions, there is no all-around guideline for


practicing engineers who are faced with the various structures
located in proximity to a ground vibration source. Hence, Table 9
was assembled considering the standard resultant. In this table max-
imum allowable PPV was classified based on four major sensitivity
degrees: highly, upper-intermediate, intermediate, and less sensitiv-
a
2.54 (0.1)

ity of various structures. Then the maximum and minimum peak


AASHTO
(1990)

allowable PPVs were obtained, which are compared in Fig. 12. It is


necessary to point out that the limits are applicable just for transient
ground vibrations, such as impact pile driving. To put the limits in a
practical way, Fig. 13 was also plotted in such a way that the struc-
tural damage level can be easily seen. Furthermore, the limits of
Telecommunication cable

considered structures in the Swedish standard are superimposed in


Commercial/industrial

Fig. 13. Eventually, it can be concluded that the minimum peak


Residential building

Underground tunnel

allowable PPV is more conservative and the maximum peak allow-


Structure
Historic–museum

Church/mosque/
Hospital–library

Electricity cable
Nuclear–marine

Gas/oil pipeline
Bridge –factory

able PPV is the limit that must be adhered to during construction


Hotel/motel

activities.
Water pipe
synagogue

Concluding Remarks

As a result of urban development, many structures are located in


Intermediate
intermediate
Sensitivity

Units are mm/s (in./s).


degree

close proximity to each other, so construction activities in the vicin-


Upper-
Highly

ity of existing structures are inevitable. On the other hand, pile driv-
Less

ing is undeniably the most prevalent technique of pile installation


due to its inherent advantages except for the vibration damage
potential to nearby structures. Therefore, the main purpose of this
Number

research is to highlight the problems associated with assessing the


ground vibrations; furthermore, numerical evaluation of the
1

© ASCE 04017029-11 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


80
Max. Peak Allowable PPV

Maximum Allowable PPV: mm/s


Min. Peak Allowable PPV
60

40

20
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

0
Less Intermediate Upper-intermediate Highly

Sensitivity Degree

Fig. 12. Maximum and minimum peak allowable PPV values of standards for transient ground vibrations

256
Max. Peak Allowable PPV
Min. Peak Allowable PPV
128 Considered Structures in Swedish Standard
Peak Particle Velocity (PPV): mm/s

64

32

16 (1)

8 (2)

(3)
4 (4)
(5)
2

1
1 2 3 4
Sensitivity Degree

Fig. 13. Structural damage level of transient ground vibrations

allowable distance from the impact pile driving site to prevent likely 3. Thus far, the determination of the maximum PPV induced dur-
structural damage to nearby buildings was performed based on the ing various construction activities has been performed using
limits of different reliable standards in terms of maximum allowable four approaches: PVS, SRSS, peak unidirectional velocity, and
PPV. Consequently, some of the prominent conclusions identified peak vertical velocity. Among the approaches peak vertical ve-
in the present investigation are as follows: locity is concluded as the best approach because it better corre-
1. To investigate the ground vibrations generated by different lates with damage and is the most prevalent and easiest
construction activities, the PPV is concluded as the best crite- approach in studies.
rion for assessing the ground vibrations. 4. Because of the implementation of a detailed modeling of the
2. It is very important to classify the ground vibrations caused by continuous impact of the pile driving process from the ground
construction activities into two types: transient vibrations (a sin- surface rather than common discontinuous pile driving meth-
gle event or multiple single events) and continuous vibrations. ods, the numerical results of verification are perfectly in accord
This is because the intensity and nature of the vibrations are with the measured data.
entirely distinct and, for this reason, standards determine differ- 5. Comparison between peak vertical and radial vibrations as
ent limits for each of the ground vibration types. vibration velocity components of the soil particles in the

© ASCE 04017029-12 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


vicinity of the impact pile driving site with the PVS of compo- Grizi, A., Athanasopoulos-Zekkos, A., and Woods, R. (2016). “Ground
nents showed that unlike the peak radial velocity, the results of vibration measurements near impact pile driving.” J. Geotech.
the peak vertical velocity are in close agreement with the PVS. Geoenviron. Eng., 10.1061/(ASCE)GT.1943-5606.0001499, 04016035.
In other words, the peak radial velocity was 40–70% smaller Hanson, C. E., Towers, D. A., Meister, L. D. (2006). “Transit noise and
vibration impact assessment.” FTA-VA-90-1003-06, Federal Transit
than the peak vertical velocity and PVS.
Administration, Office of Planning and Environment, Washington, DC.
6. The maximum allowable distance of various structures from a Head, J. M., and Jardine, F. M. (1992). “Ground-borne vibrations arising from
pile driving site was investigated in terms of the maximum piling.” Construction Industry Research and Information Association,
allowable PPV for transient vibrations existing in well-known London.
reliable standards and a report including the Swiss Standard Henke, S. (2010). “Influence of pile installation on adjacent structures.” Int.
SN640312 (SNV 1992), Eurocode 3 (CEN 1993), FHWA J. Numer. Anal. Methods Geomech., 34(11), 1191–1210.
Dynamic Compaction Report (Lukas 1995), Swedish Standard Hiller, D. M., and Hope, V. S. (1998). “Groundborne vibration generated by
SS 02 52 11 (SIS 1999), and German Standard DIN 4150-3 mechanized construction activities.” Proc. Inst. Civ. Eng. Geotech. Eng.,
(DIN 1999). It was concluded that Swedish Standard SS 02 52 131(4), 223–232.
Ho, C. E., and Tan, C. G. (2003). “Vibrations due to driving of large dis-
11 (SIS 1999) had been the most comprehensive standard avail-
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

placement concrete piles in residual granitic soils.” Proc., 6th Int. Symp.
able thus far because it considers factors affecting the ground on Field Measurements in Geomechanics, Oslo, Norway, Swets &
vibrations, such as soil type, structure type, structure materials, Zeitlinger B.V., Lisse, Netherlands, 111–116.
and foundation type. Jones and Stokes Associates. (2004). “Transportation and construction-
7. The maximum PPV and subsequently maximum allowable dis- induced vibration guidance manual.” Sacramento, CA.
tance for buried structures in the standards are definitely con- Karlsson, A. B. (2013). “Add the accident commission to investigate con-
servative because they consider the maximum PPV values at the struction missions.” hhttp://www.dn.se/debatt/tillsatt-haverikommission
ground surface over the buried structures and none of them con- -for-att-utreda-byggmisstagi (Jan. 11, 2013).
sidered the reduction due to locating in different ground depths. Kennedy, B. A. (1990). Surface mining, 2nd Ed., Society for Mining,
Metallurgy, and Exploration, Littleton, CO.
8. For practical applications, the maximum and minimum peak
Liden, M. (2012). “Ground vibrations due to vibratory sheet pile driving.”
allowable PPVs considering sensitivity degrees were presented M.Sc. thesis, Royal Institute of Technology, Stockholm, Sweden.
based on the standard resultant. The maximum allowable PPV Liyanapathirana, D. S., and Ekanayake, S. D. (2016). “Application of EPS
must be adhered to so structural damage is prevented. geofoam in attenuating ground vibrations during vibratory pile driving.”
Geotext. Geomembr., 44(1), 59–69.
Lukas, R. G. (1995). “Geotechnical engineering circular no. 1: Dynamic
References compaction.” FHWA-SA-95-037, Federal Highway Administration,
Washington, DC.
AASHTO. (1990). “Standard recommended practice for evaluation of Lupiezowiec, M., Pradelok, S., Betkowski, P., and Poprawa, G. (2014).
transportation-related earthborne vibrations.” Washington DC. “FEM model of vibration propagation in the soil caused by prefabricated
Abaqus 6-14-3 [Computer software]. SIMULIA, Providence, RI. driven piles.” Proc., 14th SGEM GeoConference on Science and
Abdel-Rasoul, E. I., and Mohamed, M. T. (2006). “Measurement and evalu- Technologies in Geology, Exploration and Mining, Albena, Bulgaria,
ation of blasting ground vibrations and airblasts at the limestone quarries 2(1), 363–368.
of Assiut Cement Company (CEMEX).” J. Eng. Sci. Assiut Univ., Mabsout, M., and Tassoulas, J. L. (1994). “A finite element model for the
34(4), 1293–1309. simulation of pile driving.” Int. J. Numer. Methods Eng., 37(2),
Ahlquist, A., and Enggren, E. (2006). “Impact on surrounding environment 257–278.
from vibro driven sheet piles.” M.Sc. thesis, Division of Soil and Rock Madheswaran, C. K., Natarajan, K., Sundaravadivelu, R., and Boominathan,
Mechanics, Royal Institute of Technology, Stockholm, Sweden. A. (2009). “Effect of open or concrete-infilled trenches on screening of
Athanasopoulos, G. A., and Pelekis, P. C. (2000). “Ground vibrations from ground vibration during pile driving.” Exp. Tech., 33(2), 43–51.
sheetpile driving in urban environment: Measurement, analysis and Martin, D. J. (1980). “Ground vibrations from impact pile driving during
effects on buildings and occupants.” Soil Dyn. Earthquake Eng., 19(5), road construction.” Supplementary Rep. 544, Transport and Research
371–387. Road Laboratory, Crowthorne, U.K.
Athanasopoulos, G. A., Woods, R. D., and Grizi, A. (2013). “Effect of pile- Masoumi, H. R., François, S., and Degrande, G. (2009). “A non-linear
driving induced vibrations on nearby structures and other assets.” Final coupled finite element-boundary element model for the prediction of
Rep. ORBP Number OR10-046, Michigan DOT, Lansing, MI. vibrations due to vibratory and impact pile driving.” Int. J. Numer. Anal.
Attewell, P. B., and Farmer, I. W. (1973). “Attenuation of ground vibrations Methods Geomech., 33(2), 245–274.
from pile driving.” Ground Eng., 6(4), 26–29. Massarsch, K. R., and Broms, B. B. (1991). “Damage criteria for small am-
BSI (British Standards Institution). (1993). “Evaluation and measurement plitude ground vibrations.” Proc., 2nd Int. Conf. on Recent Advances in
for vibration in buildings. Part 2: Guide to damage levels from ground- Geotechnical and Earthquake Engineering and Soil Dynamics,
bourne vibration.” BS 7385-2:1993, London. Missouri Univ. of Science and Technology, St. Louis, 2, 1451–1459.
Çelebi, E., Fırat, S., Beyhan, G., Çankaya, I., _ Vural, I.,_ and Kırtel, O. Massarsch, K. R., and Fellenius, B. H. (2015). “Engineering assessment of
(2009). “Field experiments on wave propagation and vibration isolation ground vibrations caused by impact pile driving.” Geotech. Eng. J.
by using wave barriers.” Soil Dyn. Earthquake Eng., 29(5), 824–833. SEAGS AGSSEA, 46(2), 54–63.
CEN (European Committee for Standardization). (1993). “Design of steel New, B. M. (1986). “Ground vibrations caused by civil engineering works.”
structures: Part 5: Piling.” Eurocode 3, ENV 1993-5, Brussels Belgium. Research Rep. No. 53, Transport and Research Road Laboratory,
Deeks, A. J., and Randolph, M. F. (1993). “Analytical modelling of hammer Crowthorne, U.K.
impact for pile driving.” Int. J. Numer. Anal. Methods Geomech., 17(5), Pourjenabi, M., and Hamidi, A. (2015). “Numerical modeling of dynamic
279–302. compaction process in dry sands considering critical distance from adja-
Dijkstra, J., Broere, W., and Heeres, O. M. (2011). “Numerical simulation cent structures.” Struct. Eng. Mech., 56(1), 49–65.
of pile installation.” Comput. Geotech., 38(5), 612–622. Ramshaw, C. L. (2002). “Computation of ground waves from pile driv-
DIN (German Institute for Standardization). (1999). “Structural vibration– ing and their effects on structures.” Ph.D. thesis, Durham Univ.,
Part 3: Effects of vibration on structures.” DIN 4150-3, Berlin. Durham, U.K.
Ekanayake, S. D., Liyanapathirana, D. S., and Leo, C. J. (2013). “Influence Rezaei, M., Hamidi, A., and Farshi Homayoun Rooz, A. (2016).
zone around a closed-ended pile during vibratory driving.” Soil Dyn. “Investigation of peak particle velocity variations during impact pile
Earthquake Eng., 53(Oct), 26–36. driving process.” Civ. Eng. Infrastruct. J., 49(1), 59–69.

© ASCE 04017029-13 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029


Rockhill, D. J., Bolton, M. D., and White, D. J. (2003). “Ground-borne Sylvestre-Williams, N. (2011). “The city of Toronto construction
vibrations due to press-in piling operations.” Proc., BGA Int. Conf. on vibration by-law: An overview and discussion.” Proc., Noise-Con
Foundations: Innovations, Observations, Design and Practice, Thomas 11, Institute of Noise Control Engineering and Transportation
Telford Limited, London, 743–756. Research Board Committee, Portland, OR, INCE-USA, Reston,
Schumann, B., and Grabe, J. (2011). “FE-based modelling of pile driving in VA, 119–124.
saturated soils.” Proc., 8th Int. Conf. on Structural Dynamics, Thandavamoorthy, T. S. (2004). “Piling in fine and medium sand–A case
EURODYN 2011, European Association for Structural Dynamics study of ground and pile vibration.” Soil Dyn. Earthquake Eng., 24(4),
(EASD), K.U.Leuven, and Technological Institute of the Royal Flemish 295–304.
Society of Engineers, Leuven, Belgium. Uromeihy, A. (1990). “Ground vibration measurements with special
Serdaroglu, M. S. (2010). “Nonlinear analysis of pile driving and ground reference to pile driving.” Ph.D. thesis, Durham Univ., Durham,
vibrations in saturated cohesive soils using the finite element method.” U.K.
Ph.D. thesis, Univ. of Iowa, Iowa City, IA. Whenham, V., Areias, L., Rocher-Lacoste, F., Vie, D., Bourdouxhe, M. P.,
SIS (Swedish Institute for Standards). (1999). “Vibration and shock— and Holeyman, A. (2009). “Full scale sheet pile vibro-driving tests.”
Guidance levels and measuring of vibrations in buildings originating Proc., 17th Int. Conf. on Soil Mechanics and Geotechnical Engineering,
from piling, sheet piling, excavating and packing to estimate permitted Alexandria, Egypt, IOS Press, 1354–1357.
Downloaded from ascelibrary.org by University of Manitoba on 02/11/21. Copyright ASCE. For personal use only; all rights reserved.

vibration levels (in Swedish).” SS 02 52 11, Stockholm, Sweden. White, D., Finlay, T., Bolton, M., and Bearss, G. (2002). “Press-in piling:
SNV (Swiss Association for Standardization). (1992). “Vibrations— Ground vibration and noise during pile installation.” International Deep
Vibration effects in buildings.” SN640312, Winterthur, Switzerland. Foundations Congress 2002, ASCE, Reston VA, 363–371.
Svinkin, M. (2004). “Minimizing construction vibration effects.” Pract. Wiss, J. F. (1981). “Construction vibrations: State-of-the-art.” J. Geotech.
Period. Struct. Des. Constr., 10.1061/(ASCE)1084-0680(2004)9: Eng. Div., 107(2), 167–181.
2(108), 108–115. Woods, R. D. (1997). “Dynamic effects of pile installations on adja-
Svinkin, M. (2006). “Mitigation of soil movements from pile driving.” Pract. cent structures.” NCHRP Synthesis 253, National Academy Press,
Period. Struct. Des. Constr., 10.1061/(ASCE)1084-0680(2006)11:2(80), Washington, DC.
80–85. Zhang, M., Tao, M., Gautreau, G., and Zhang, Z. (2013). “Statistical
Svinkin, M. (2015). “Tolerable limits of construction vibrations.” Pract. approach to determining ground vibration monitoring distance during
Period. Struct. Des. Constr., 10.1061/(ASCE)SC.1943-5576.0000223, pile driving.” Pract. Period. Struct. Des. Constr., 10.1061/(ASCE)SC
04014028. .1943-5576.0000156, 196–204.

© ASCE 04017029-14 Pract. Period. Struct. Des. Constr.

Pract. Period. Struct. Des. Constr., 2018, 23(1): 04017029

You might also like