Hope, V. S., & Hiller, D. M. (2000) - The Prediction of Groundborne Vibration From Percussive Piling

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Color profile: Generic CMYK printer profile

Composite Default screen

700

NOTES

The prediction of groundborne vibration from


percussive piling
V.S. Hope and D.M. Hiller

Abstract: Environmental assessments of proposed infrastructure projects include an appraisal of the vibration that
would be generated by potential construction methods. This paper considers one part of that process: prediction of
groundborne vibration from percussive piling. Various methods for predicting the magnitude of the groundborne vibra-
tion generated by percussive piling have been described in the literature. In general, the accuracy of the predictors is
limited. The paper investigates the reasons for this, through the development of theoretical models and analysis of field
data from an extensive programme of on-site vibration monitoring. It is found that some widespread assumptions about
the relationship between the energy rating of a percussive piling hammer and the vibrational energy developed in the
ground are invalid. Ground conditions are shown to have a dominant influence on the magnitude of the groundborne
vibration generated by percussive piling, and piling vibration predictors which take no account of soil type yield con-
siderably less accurate estimates than site-specific predictions.

Key words: piling, vibration, monitoring, prediction, environmental assessment.

Résumé : Les évaluations environnementales de projets d’infrastructure comprennent une estimation de la vibration qui
serait générée par les méthodes potentielles de construction. Cet article examine une partie de ce processus : la
prédiction de la vibration transmise par le sol durant le fonçage de pieux par percussion. Diverses méthodes pour
prédire la magnitude de la vibration transmise par le sol et générée par le fonçage de pieux par percussion ont été
décrites dans la littérature. En général, la précision des méthodes de prédiction est limitée. Cet article en étudie les
raisons au moyen du développement de modèles théoriques et de l’analyse de données de terrain provenant d’un vaste
programme de mesures de vibrations sur place. On trouve que des hypothèses très répandues sur la relation entre
l’estimation de l’énergie d’un marteau de percussion pour les pieux et l’énergie de vibration développée dans le sol
sont invalides. On montre que les conditions du sol ont une influence dominante sur la magnitude de la vibration
générée par le fonçage de pieux par percussion et transmise par le sol, et que les méthodes de prédiction de ces
vibrations qui ne prennent pas en compte le type de sol donnent des estimations considérablement moins précises que
les prédictions spécifiques au site.

Mots clés : fonçage de pieux, vibration, mesure, prédiction, évaluation environnementale.

[Traduit par la Rédaction] Notes 711

Introduction Act 1990). The accurate estimation of likely vibration levels


at a site is important (Hiller and Hope 1998); however, the
The groundborne vibration generated by percussive piling reliability of empirical piling vibration predictors is variable.
operations is usually insufficient to cause damage to nearby
The research presented here forms parts of a larger project
structures (BSI 1992a) but it may be perceptible to people
to investigate groundborne vibration arising from construction
living and working around a site (BSI 1992b) or cause con-
operations. At the outset of that project, it was assumed that
cern for vibration-sensitive installations and fragile struc-
the ground vibration effects of percussive piling are relatively
tures. Environmental impact assessments of construction
well understood. However, through analysis of the field data
projects include a quantitative estimate of the vibration that
obtained, it became apparent that several questions are still to
may be caused (EC 1985) and, in the U.K., local authorities
be resolved. The aim of this paper is to set out these issues.
can suspend works if they cause intrusion “so as to be preju-
The paper considers the factors that have been suggested to
dicial to health or a nuisance” (Environmental Protection
influence the magnitude of ground vibration from percussive
piling and assesses their relevance in the predictive modelling
Received March 15, 1998. Accepted November 12, 1999. of piling-induced vibration using field data and theoretical
V.S. Hope. Department of Civil Engineering, University of models. It is demonstrated that the widely assumed depend-
Surrey, Guildford, GU2 7XH, U.K. ence of ground vibration magnitude on hammer impact en-
D.M. Hiller. Transport Research Laboratory, Crowthorne, ergy is much less strong than might be expected from its
Berkshire, RG45 6AU, U.K. central role in empirical predictors. The magnitude of piling-

Can. Geotech. J. 37: 700–711 (2000) © 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:30:34 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Notes 701

Table 1. Values of parameters used to generate the prediction curves in Fig. 1.

Predictor Parameters Data fitting


Attewell and Farmer 1973 x = 1, k = 1.5 Upper bound
Whyley and Sarsby 1992 x = 1, k = 0.25 for soft loose soil; k = 0.75 for stiff or Upper bound
medium dense soil; k = 1.5 for stiff or dense soil
Attewell et al. 1992a, 1992b k1 = –0.073, k2 = 1.38, k3 = 0.234 1 SD above mean
Hiller and Crabb 1998 k = 3 for stiff or medium dense soil Upper bound
Lo 1977 α = 0.013 m–1, vo = 80 mm·s–1, r0 = 1 m Deepest penetration
Head and Jardine 1992 x = 1, k = 1.5 (for r > 0.5 m) Upper bound
BSI 1992a x = 1, k = 0.75 Upper bound
CEN 1998 x = 1, k = 1.0 for very stiff cohesive soil; k = 0.75 for Upper bound
stiff cohesive soil; k = 0.5 for soft cohesive soil

induced vibration is shown to be highly influenced by ground  W  W


conditions, and it is recommended that predictions of [2] log v res = k1 + k2 log   + k3 log2  
groundborne vibration from percussive piling should take ac-  r   r 
count of both the soil type and the driver energy.
where vres is the peak resultant particle velocity; and values
for the constants of proportionality k1, k2, and k3 for percus-
Published prediction methods sive piling were determined empirically.
Strategies for predicting vibration magnitudes for piling Van Staalduinen and Waarts (1992) developed a predictor
have taken two forms: (i) empirical predictions, developed using the attenuation formulation proposed by Mintrop
using field data from many sites; and (ii) pilot-scale trials at (cited by Bornitz 1931) for Rayleigh surface waves:
the site of interest (for example, Grose 1986). One of the
r0
earliest empirical predictors was developed by Attewell and [3] vr = v 0 e − α (r −r 0 )
Farmer (1973), who analysed piling vibration data from sev- r
eral sites and suggested that
where vr is the peak vibration at distance r ;
x
 W v0 is the peak vibration at distance r0;
[1] vv = k   r is the radial distance between the pile and the monitor-
 r  ing point;
r0 is the radial distance from the pile to a point of known
where vv is the peak vertical particle velocity (in mm/s); vibration magnitude; and
W is the input energy (in J), assumed to be the hammer α is a material damping coefficient (of dimension length–1).
impact energy; Van Staalduinen and Waarts found that v0 correlated with
r is the radial distance between the pile and the monitor- cone penetration (CPT) resistance and used CPT data to pro-
ing point (m); vide a degree of site specificity to the prediction at less cost
x is an empirically determined index; and than a full piling trial. Their method was applied at one site
k is an empirically determined (dimensional) constant of only so its general validity was not tested. Lo (1977) based a
proportionality. vibration prediction method on eq. [3] in which values of v0
Predictors having the general form of eq. [1] have been and α were obtained from piling trials.
widely adopted (for example, in BSI 1992a; Head and Jar- Equations [1] and [3] differ fundamentally in their treat-
dine 1992; Hiller and Crabb 1998), although both BSI ment of attenuation. The attenuation model in eq. [1] de-
(1992a) and Head and Jardine (1992) recommended that scribes, for x of unity (Attewell and Farmer 1973),
close to a pile the slope distance from the pile toe, s, is used geometric spreading of body waves. In contrast, eq. [3] mod-
rather than the radial distance r. Whyley and Sarsby (1992) els material damping and the geometric spreading of surface
and Hiller and Crabb (1998) reported, as Wiss (1967) had waves.
also shown, that k varies with geology. Whyley and Sarsby Figure 1 compares several of the published vibration pre-
assigned values to k ranging from 0.25 to 1.5, depending on dictors with ground vibration data from percussive piling op-
soil type. A similar approach has been adopted in the Euro- erations at a site monitored in the present study. Table 1
pean code (CEN 1998). summarises the parameters used in each predictor (eq. [3]
Using field data presented by Uromeihy (1990), Attewell was implemented in the manner suggested by Lo 1977, and
et al. (1991) observed that maxima occurred in vibration – α was determined from a subset of the field data). The pre-
radial distance data plots at about r = 10 m. Figure 1a shows dictors presented in Fig. 1 use differing models and parame-
this effect, which has been attributed to superposition of ters. To establish effective prediction strategies, it is
body and surface waves emanating from the pile toe and the necessary to clarify which processes are in operation and
shaft, respectively (Jongmans 1996). To reduce overconser- which parameters are important. The following sections
vatism of predictions at close range, Attewell et al. (1992a, present an examination of the physical basis of possible
1992b) recommended that a quadratic relationship be used, parameterizations of the piling vibration problem and the re-
specifically sults of analyses of ground vibration data from percussive

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:30:35 PM
Color profile: Generic CMYK printer profile
Composite Default screen

702 Can. Geotech. J. Vol. 37, 2000

Fig. 1. A posteriori predictions, obtained using published predictors, of the groundborne vibration from percussive piling compared
with field data from site A plotted against (a) horizontal radial distance, and (b) slope distance.

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:30:43 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Notes 703

Table 2. Typical acoustic impedance values for pile and ground materials and the corresponding energy transmission coefficients for
normal incidence (eq. [5]).

Mass density P-wave speed, Acoustic impedance, Transmission coefficient, T


Material ρ (kg·m–3) a (m·s–1) ρa (Z) (MPa·s·m–1) Steel Concrete
Steel 7800 5700 44.5 — —
Concrete 2400 5000 12.0 — —
Sand (saturated) 2000 1500 3.0 0.24 0.64
Clay (stiff) 2300 2000 4.6 0.34 0.80
Sandstone 2400 2300 5.5 0.39 0.86

Fig. 2. Schematic of vibrational wave fronts radiating from a centred on the longitudinal axis of the pile (Fig. 2a). For
percussive driven pile: (a) shaft source; (b) toe source. such wave fronts, the rate of attenuation of wave energy in-
tensity, due to geometric spreading, would be proportional
to 1/r, where r is the horizontal radial distance from the pile.
Since the peak particle velocity (ppv) of a wave is propor-
tional to the square root of the energy of the wave, the ppv
would diminish as 1/(r)0.5.
For the frictional pile model, the energy imparted from the
pile to the ground would depend on the shear force acting
along the pile shaft and the distance through which this
force moves at each impact. The portion of this energy asso-
ciated with elastic ground deformation is the groundborne
vibrational energy, Wf, per blow, given by
[4] Wf ∝ Ff d
where Ff is the resistive shear force along the pile shaft, and
d is the distance moved by the pile (the “set”) per blow. The
force Ff will vary with the shear stress mobilised at the
shaft–soil interface (Billet and Sieffert 1989; Coop and
Wroth 1989) and, for a given shear stress, will increase with
increasing pile embedment length, l. The set, d, will depend
on the hammer energy, WH, and on the properties of the
ground penetrated, and therefore these factors will also in-
fluence Wf.
In the case of a wholly “end-bearing” pile, with no shaft
piling in terms of these parameters. Figure 1 is reconsidered
resistance effective during driving, the compressive stress
in more detail in the Discussion.
pulse initiated by the impact will propagate down the shaft
to the pile toe (Fig. 2b). Transmission of energy across the
Conceptual models and parameterization pile–soil interface will occur, governed by the contrast in
In foundation design, an axially loaded pile can be consid- acoustic impedance properties of the pile and soil. Spe-
ered in terms of a normal force acting on the pile base and cifically, the ratio of the energy intensities of the transmitted
shear forces along the pile shaft. This elementary representa- and incident waves, T, is given by (Telford et al. 1976;
tion is used here to develop a conceptual model of how a Zoeppritz 1919)
blow to the top of a partially embedded, impact-driven pile 4Z pile Z soil
may give rise to groundborne vibration during pile installation. [5] T =
The magnitude of the vibration depends on the amount of (Z pile + Z soil ) 2
energy imparted to the ground at the source, the rate of
attenuation of the energy as it propagates through the for normal incidence at the boundary, where the acoustic im-
ground, and the distance over which this attenuation occurs. pedance, Z, of a material is the product of its mass density
The development of the model is therefore treated in three and compressional wave propagation speed (Table 2 presents
parts: input energy, attenuation rate, and attenuation distance typical values for Z and T ). In the absence of any other ef-
between source and receptor. fect, the energy entering the ground, Web, from an end-
In the idealised case of a “frictional” pile, with no end re- bearing pile would be
straint during driving, the stress pulse induced by the ham- [6] Web ∝ TWH
mer impact propagates along the shaft and is reflected at the
toe. Resistive shear forces on the partially embedded shaft However, part of WH is used to drive the pile; this is consid-
would induce vertically polarised shear waves in the sur- ered further in the Discussion.
rounding ground. These would propagate outwards as cylin- In the end-bearing paradigm, the toe is the point at which
drical or conical (see Attewell and Farmer 1973) wave fronts energy enters the ground, and therefore the slope distance, s,

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:30:44 PM
Color profile: Generic CMYK printer profile
Composite Default screen

704 Can. Geotech. J. Vol. 37, 2000

Fig. 3. Near-surface geology at site A, determined at two bore- which necessarily restricted the scope of the investigations
holes adjacent to the piling area. that could be conducted and the experimental control that
could be achieved. However, the range of distances over
which data were acquired and the duration of the data re-
cords were unusually extensive. Three-component measure-
ments of particle velocity were recorded for the full period
of driving, using surface triaxial geophone clusters ranged
between a few metres to about 100 m from each pile.

Distance
The first parameter to be examined is the distance over
which attenuation occurs. For this, data are considered from
site A, at which the same steel H-piles and piling rig were
used for all operations. Figure 3 shows two borehole logs
from site A. Figure 4a presents the peak resultant particle
velocity, vres, plotted against the horizontal distance, r, from
the pile. Figure 4b presents the same data plotted against the
slope distance, s, from the pile toe to each monitoring point
on the ground surface. In Fig. 4a, the relationship between
ppv and r is not linear near the pile. In Fig. 4b, the distribu-
tion of the data is more closely linear, and can be character-
ised as
1
[7] vres ∝
sn
where s is the slope distance between the pile tip and the
monitoring point;
vres is the peak resultant particle velocity; and
n is an attenuation index.
For these data, n = 1.2 with a correlation coefficient of
–0.78. Equation [7] can be applied throughout the range of
distances for which data are available from site A. The use
of s rather than r therefore appears preferable.
between the toe and a monitoring point (Wiss 1967; BSI
1992a) would be relevant when considering attenuation. At Hammer impact energy
depth, a pile toe will act as a point source of body waves From elastic theory, a square-root relationship between
having spherically expanding wave fronts (Fig. 2b) for the peak particle velocity (ppv) and energy of a wave is ex-
which, in the absence of material damping, the ppv will di- pected. The energy term, W, in eq. [1] and its variants are
minish as 1/s. At the ground surface, the particle velocity of usually taken to be the energy rating of the hammer, WH
body waves diminishes as 1/s2 (Ewing et al. 1957; Lo 1977). (Wiss 1967; Attewell and Farmer 1973; Lo 1977; Whyley
In practice, both pile-driving models may operate. In clayey and Sarsby 1992; CEN 1998). To investigate this assump-
soils, an increase in pore-water pressure and remoulding of tion, an experiment was conducted at a second site (site B)
the soil around the pile shaft occur during driving (Poulos in which hammer drop heights, h, of 1.0, 0.5, 0.25, 0.1, and
and Davis 1980; Tomlinson 1980; Coop and Wroth 1989) then 1.0 m again were used sequentially. Figure 5 presents
and the shear stress mobilised at the soil–shaft interface is the results obtained, including the observed penetration per
consequently reduced. In granular soils, the displacement of blow (set). In Fig. 5b, the ground vibration data are normal-
soil around an advancing pile may increase the skin friction ized by the ppv from the final 1.0 m drop. From Fig. 5b, it
on the shaft (Tomlinson 1980; Poulos and Davis 1980). can be suggested that the magnitude of the groundborne vi-
Therefore, in contrast with typical post-installation behav- bration generated by percussive piling is related to the po-
iour of piled foundations, during installation a pile driven tential energy of the hammer mechanism, WH. However, in
into clay may display the end-bearing model of ground vi- Fig. 5b the relationship between vres and h0.5, and hence
bration generation (i.e., a toe source), whereas in sand or (WH)0.5, is not linear. This is in contrast with the assump-
gravel a pile may show frictional vibration generation (i.e., a tions made in vibration prediction formulae (see eqs. [1] and
shaft source), or aspects of both behaviour. [2]). As well as ppv, Fig. 5b shows that the set achieved per
blow, d, also increased nonlinearly with increasing hammer
Analysis of field measurements energy. The possible connections between hammer energy,
pile driving energy, and groundborne vibration energy are
Measurements of groundborne vibration from percussive considered in the Discussion.
piling operations have been made by the second author at Figure 6a shows ppv and hammer energy data from sev-
several sites in the United Kingdom. Access to the sites was eral sites and piling rigs. It includes data published by
predicated on not interrupting construction progress on site, Uromeihy (1990), to extend the data range of the present

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:02 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Notes 705

Fig. 4. Groundborne vibration data from percussive piling at site A, grouped by pile toe depth: (a) peak resultant particle velocity
against horizontal radial distance; (b) peak resultant particle velocity against slope distance. A Banult 700 piling rig with a 5 t drop
hammer was used to drive steel H-section piles (UBP 305 × 305 × 186 kg/m).

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:05 PM
Color profile: Generic CMYK printer profile
Composite Default screen

706 Can. Geotech. J. Vol. 37, 2000

Fig. 5. Variation of peak resultant particle velocity and pile set with hammer drop height, h, at site B: (a) unadjusted data; (b) ppv
normalised by final impact from h = 1 m, with set data.

study. In contrast with the results from site B (Fig. 5), no and hammer energy is unlikely to yield reliable predictions
strong dependence of ppv on WH is discernible in Fig. 6a. of groundborne vibration from percussive piling.
However, the data in Fig. 6a are from pile–receptor dis-
tances ranging from about 2 to 100 m (see Fig. 6b). To at- Depth of embedment
tempt to eliminate the effect of distance from the data, in In Fig. 4, the field data from site A are classified by the
Fig. 6c the hammer energy has been normalized by the slope depth of embedment of the pile. There is a trend of increas-
distance, s, raised to the power n (eq. [7]), where a value of ing ppv with depth. Figure 7 presents data from site A re-
n = 1.1 was obtained from regression analysis of the data in corded at the monitoring station farthest from the piles
Fig. 6b (the reciprocal of the distance-normalized hammer (about 100 m distant). At this offset, the slope distance be-
energy ((WH)0.5/s1.1) is used in Fig. 6c to facilitate visual tween the pile toe and the monitoring station increased by
comparison with Fig. 6b). The scatter in Fig. 6c extends at only 4% during driving. Despite the negligible change in at-
least an order of magnitude, and is almost two orders of tenuation distance (see eq. [7]), there is almost an order of
magnitude if outliers are considered. This suggests that an magnitude rise in ppv during driving. According to eq. [1], a
empirical predictive model which incorporates only distance 100-fold increase in W would be required to achieve this.

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:07 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Notes 707

Fig. 6. Field data from several sites and piling rigs, shown as peak resultant particle velocity versus (a) hammer energy, (b) slope dis-
tance, and (c) the square root of the hammer energy adjusted for slope distance.

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:31 PM
Color profile: Generic CMYK printer profile
Composite Default screen

708 Can. Geotech. J. Vol. 37, 2000

Fig. 6 (concluded).

However, at site A, aside from “toeing in” the first 4 m of Fig. 7. Variation of peak resultant particle velocity with depth of
shaft, the same hammer energy, WH, was used throughout embedment of pile at 100 m radial distance from percussion-
driving. Figures 4, 6c, and 7 suggest that the groundborne driven piles at site A.
energy generated by percussive piling is more strongly influ-
enced by a factor which varies with depth of penetration
than by WH. This additional factor may explain the apparent
contradiction within the field results in Figs. 5b and 6a con-
cerning the influence of WH on ppv, since the measurements
presented in Fig. 5b were obtained from a pile as it was
driven through about 0.1 m only, whereas for Fig. 6a moni-
toring continued from toeing in to full installation.

Ground conditions
As a pile is driven (i) the length of shaft in contact with
the surrounding soil increases, (ii) the properties of the soil
encountered by the shaft may vary due to heterogeneity, and
(iii) the soil encountered by the pile toe may vary. Each of
these factors could account for the behaviour shown in, for
example, Fig. 7. The data in Fig. 7 can be interpreted as
indicating either that the ppv increases with increasing
embedment length, l, or that the ppv increases as the toe en-
counters different strata. That is, it is difficult to distinguish
between effects at the shaft and at the pile toe.
To attempt to clarify the ambiguity, other data from site A
have been examined. Figure 8 presents the driving record of
a pile at the site (expressed as the number of impact blows the variation of the standard penetration test (SPT) N count
required to drive a pile each additional 250 mm, plotted with depth at two test locations near the pile. In its operation
against the current total depth of penetration). Such records (BSI 1990; Clayton 1995), the SPT apparatus is akin to an
indicate the resistance of the ground to pile penetration. This impact-driven pile that is isolated from the surrounding soil
resistance may be a compound of end resistance and shaft along its shaft. In Fig. 8, the SPT N count and ppv values
friction, or one factor may dominate. Figure 8 also shows both begin to increase at depths below about 12–14 m,

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:37 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Notes 709

Fig. 8. Pile driving record from site A compared with nearby SPT N values, and peak resultant particle velocity data.

which corresponds with the onset of denser material under- a short distance with that needed to overcome a lesser force
lying the soft alluvial deposits at the site (Fig. 3). Both the which is acting over a greater distance.
SPT and ppv values increase further, and the pile penetration The second explanation presented in the literature consid-
resistance starts to increase, as the pile begins to penetrate ers the partition of energy at the pile–soil interface (Attewell
the stiff Kimmeridge clay, from about 16 m. From these ob- and Farmer 1973). Since displacement piles are stiffer (and
servations, it can be concluded that, at site A, the magnitude hence have greater acoustic impedance) than the ground into
of the groundborne vibration generated by percussive piling which they are driven, the energy transmission ratio at the
is affected more by the soil encountered at the pile toe than pile–soil interface will increase as the acoustic impedance of
that along the shaft. This conclusion is supported by the ear- the ground increases which, in general, occurs as the ground
lier observation that the geometric attenuation of the stiffness increases (Table 2). With reference to Table 2, for a
groundborne energy at site A varies with the distance, s, given pile material, the range of Zsoil values that may be en-
from the pile toe rather than the distance, r, from the shaft. It countered during driving could account for a 30–60% varia-
should be noted that, although data from only one site are tion in T, and hence in ppv (eqs. [5] and [6]). This limited
presented here, monitoring of percussive piling operations at range contrasts with the almost 10-fold variation of ppv with
other sites also indicated a similar general trend of increas- depth shown in Fig. 7 and suggests that acoustic impedance
ing ppv with increasing pile toe depth. effects cannot alone account for observed field behaviour.
It appears that each explanation has some validity, but
some flaws also. Understanding which explanation, if either,
Discussion is correct is significant, since they lead to divergent predic-
Groundborne vibrations from percussive piling are often tions in some situations. For example, for two piles made of
observed to be greater in stiff, dense soils than in loose, soft differing materials but which have the same dimensions and
soils (refer, for example, to Whyley and Sarsby 1992; are driven with equal hammer energies into the same
Ciesielski et al. 1980; Hosking et al. 1988; Heckman and ground, the difference in the acoustic impedance of the piles
Hagerty 1978; Lo 1977; d’Appolonia 1971; Wiss 1967; would be expected to give rise to differing levels of ground
Luna 1967; Baba and Toriuim 1957). Two separate explana- vibration in each case. In contrast, the resistance to driving
tions for this behaviour have been proposed in the literature. and therefore, according to d’Appolonia (1971), the ground
D’Appolonia (1971), amongst others, suggested that resis- vibration generated would be expected to be similar for the
tance to pile penetration governs how much, after losses (see two piles, since they have the same dimensions.
Wiss 1967), of the available hammer energy WH goes into It is plausible that both mechanisms operate. Perhaps a
the work of driving the pile (that is, permanent plastic defor- proportion (governed by the transmission ratio T (eq. [5])) of
mation of the ground) and how much is available to become the hammer energy, WH, is transmitted directly from the pile
groundborne energy (that is, transient elastic deformation). toe to the soil as groundborne vibrational energy and, ne-
In resistive soils, it was suggested, the set is low and signifi- glecting losses, the remaining energy is available to drive the
cant groundborne vibration is therefore produced, but in eas- pile. Of this energy, some will cause elastic ground deforma-
ily penetrated soils most of the hammer energy goes to tion. It is notable that the soil properties relevant to each vi-
advance the pile and vibration is consequently low. A diffi- bration generation mechanism are different: the transmission
culty of this explanation is that it attempts to compare the ratio, T, depends on the stiffness and density of the ground,
work needed to overcome a large resistive force acting over whereas pile penetration depends on soil strength.

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:39 PM
Color profile: Generic CMYK printer profile
Composite Default screen

710 Can. Geotech. J. Vol. 37, 2000

Figure 1 compared several predictors and their perfor- pile, the magnitude of groundborne vibration varies signifi-
mance against field data. In Fig. 1a, most of the prediction cantly with soil type and therefore, in heterogenous ground,
lines lie towards the upper edge of the data spread. Beyond with depth of penetration also. A predictor that does not take
about r = 11 m, these lines follow the peak values closely account of soil type cannot yield accurate predictions of pil-
and discrepancies between the predictors are slight. At close ing-induced groundborne vibration. If full trials cannot be
ranges (r ≤ 11 m in Fig. 1a), there is considerable over- conducted at a site, then published soil-dependent k factors
prediction of the vibration magnitudes. The overprediction is should be utilised (such as those given by Whyley and
significant when judged against the threshold levels for tran- Sarsby 1992). In variable ground, the largest soil-dependent
sient vibration which may cause cosmetic damage to light- k factor for the strata encountered should be used.
framed structures (for example, BSI 1993 gives threshold
levels of 20–50 mm/s, depending on frequency). At close
Acknowledgement
range, the predicted vibration values in Fig. 1a exceed the
threshold levels for cosmetic damage, whereas the observed Field data were acquired by the Transport Research Labo-
groundborne vibrations at site A were well below the thresh- ratory through funding from the United Kingdom Depart-
old. Figure 1b presents predictions which are based on use ment of the Environment, Transport and the Regions. Views
of the slope distance, s. For these, the problem of over- expressed do not reflect those of either organization.
prediction at close range is less marked than in Fig. 1a.
Several of the predictors in Fig. 1 were developed empiri-
cally by upper bound data fitting (see Table 1) and thus are References
intentionally conservative. Had the standard installation Attewell, P.B., and Farmer, I. 1973. Attenuation of ground vibra-
depth at site A been less than 30 m (Fig. 4), these predictors tion from pile driving. Ground Engineering, June, 1973, pp. 26–
would have overestimated the maximum groundborne vibra- 29.
tion generated at the site. Attewell, P.B., Selby, A.R., and Uromeihy, A. 1991. Non-mono-
The lines in Fig. 1 that were generated using soil- tonical decay of ground surface vibrations caused by pile driv-
dependent predictors (Whyley and Sarsby 1992; CEN 1998; ing. In Earthquake, blast and impact. Measurement and effects
Hiller and Crabb 1998) are distinct from those derived using of vibration. The Society for Earthquake and Civil Engineering
predictors in which ground conditions are not considered. Dynamics, London, U.K. pp. 463–481.
For Whyley and Sarsby (see Fig. 1a) and CEN (see Fig. 1b) Attewell, P.B., Selby, A.R., and O’Donnell, L. 1992a. Estimation
three prediction lines have been generated using the pub- of ground vibration from driven piling based on statistical analy-
lished k values for generic soil conditions (Table 1). In ses of recorded data. Geotechnical and Geological Engineering,
Fig. 1b, the prediction lines generated using the CEN predic- 10: 41–59.
Attewell, P.B., Selby, A.R., and O’Donnell, L. 1992b. Tables and
tor overestimate the lowest vibrations and underestimate the
graphs for the estimation of ground vibration from driven piling
largest vibrations encountered at site A. In Fig. 1a, the pre-
operations. Geotechnical and Geological Engineering, 10: 61–
dictions based on the predictor of Whyley and Sarsby 87.
bracket the field data: specifically, k = 0.25 (soft alluvium) Baba, Y., and Toriuim, I. 1957. Vibrations underground by pile-
provides a lower bound and k = 1.5 (stiff clay) gives an up- driving. Technical Report, Osaka University, Vol. 7, No. 239,
per bound to the field data. With reference to the site geol- pp. 83–94.
ogy (Fig. 3) and the pile toe depth data indicated in Fig. 4, it Billet, P., and Sieffert, J.G. 1989. Soil – sheet pile interaction in
can be suggested that, at a given site, the highest k value for vibro-piling. Journal of Geotechnical Engineering, ASCE,
the soils likely to be encountered during driving should be 115(8): 1085–1101.
used in predictions of the groundborne vibration from per- Bornitz, G. 1931. Über die Ausbreitung der von Groβ kolbenmaschinen
cussive piling. erzeugten Bodenschwingungen in die Tiefe. J. Springer, Berlin.
BSI. 1990. Determination of the penetration resistance using the
split-barrel sampler (the standard penetration test, SPT). BS
Conclusions 1377, Part 9, Section 3.3. British Standards Institution, London.
Accurate prediction of groundborne vibration from per- BSI. 1992a. Noise control on construction and open sites. Part 4.
cussive piling, whether using empirical predictors or pilot- Code of practice for noise and vibration control applicable to
scale field trials, requires appropriate parameterization of the piling operations. BS 5228, Part 4. British Standards Institution,
London.
problem within a tractable model. Previous work in this field
BSI. 1992b. Evaluation of human exposure to vibration in build-
has led to several predictive models which have the benefit
ings (1 Hz to 80 Hz). BS 6472. British Standards Institution,
of being easy to implement, but which do not in general give London.
reliably accurate predictions. Based on data analyses pre- BSI. 1993. Evaluation and measurement for vibration in buildings:
sented here, it is recommended that the geometric spreading Part 2. Guide to damage levels from groundborne vibration. BS
of groundborne vibration from percussive piling should be 7385, Part 4. British Standards Institution, London.
modelled using the slope distance, s, between the pile toe CEN. 1998. European Prestandard ENV 1993-5, Eurocode 3: De-
and the monitoring point rather than the horizontal radial sign of steel structures — part 5: piling. European Committee
distance, r. The rate of attenuation of ground vibration with s for Standardization (CEN), Brussels.
does not vary greatly between sites and is of the order of Ciesielski, R., Maciag, E., and Stypula, K. 1980. Ground vibra-
n = 1.1–1.2 (eq. [7]). Although the nominal energy rating of tions induced by pile driving: some results of experimental
the hammer driving mechanism serves as a starting point investigations. In Proceedings of the International Symposium
when estimating the energy that enters the ground from a on Soils under Cyclic and Transient Loading, Swansea, U.K.

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:40 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Notes 711

Edited by G.N. Pande and O.C. Zienkiewicz. A.A. Balkema, Tomlinson, M.J. 1980. Foundation design and construction. 4th ed.
Rotterdam, The Netherlands, Vol. 2, pp. 757–762 Pitman, London
Clayton, C.R.I. 1995. The standard penetration test (SPT): methods Uromeihy, A. 1990. Ground vibration measurements with special
and use. CIRIA Report 143, Construction Industry Research and reference to pile driving. Ph.D. thesis, University of Durham,
Information Association, London. Durham, U.K.
Coop, M.R., and Wroth, C.P. 1989. Field studies of an instru- van Staalduinen, P.C., and Waarts, P.H. 1992. Prediction of vibra-
mented model pile in clay. Géotechnique, 39(4): 679–696. tions due to pile driving. In Proceedings of the 4th International
D’Appolonia, D.J. 1971. Effects of foundation construction on Conference on the Application of Stress-Wave Theory to Piles.
nearby structures. In Proceedings of the 4th Pan-American Con- A.A. Balkema, Rotterdam, The Netherlands, pp. 159–166.
ference on Soil Mechanics and Foundation Engineering, Vol. 1, Whyley, P.J., and Sarsby, R.W. 1992. Groundborne vibration from
pp. 189–236. piling. Ground Engineering, May, 1992, pp. 32–37.
EC. 1985. Council Directive of 27 June 1985 on the assessment of Wiss, J.F. 1967. Damage effects of pile driving vibration. Highway
the effects of certain public and private projects on the environ- Research Record, 155: 14–20.
ment (85/337/EEC). Official Journal of the European Commu- Zoeppritz, K. 1919. Über reflexion und durchgang seismischer
nities, No. L175, 5-7-85, pp. 40–48 wellen durch Unstetigkerlsfalschen. Über Erdbebenwekken VII-
Environmental Protection Act. 1990. Chapter 43. Her Majesty’s B, Nachrichten der Königlichen Gesellschaft der Wissenschafen
Stationery Office, London. zu Göttingen, math.-phys. K1, pp. 57–84. Berlin.
Ewing, W.M., Jardetzky, W.S., and Press, F. 1957. Elastic waves in
layered media. McGraw-Hill Book Company, New York.
Grose, W.J. 1986. Driving piles adjacent to vibration-sensitive List of symbols
structures. Ground Engineering, May, 1986, pp. 23–31.
Head, J.M., and Jardine, F.M. 1992. Ground-borne vibrations aris- a compressional wave propagation speed
ing from piling. CIRIA Technical Note 142, Construction Indus- d pile set, i.e., permanent pile penetration achieved per
try Research and Information Association, London. blow
Heckman, W.S., and Hagerty, D.J. 1978. Vibrations associated with Ff total resistive force along pile shaft
pile driving. Journal of the Construction Division, ASCE, h hammer drop height
104(C04): 385–394. k constant of proportionality (eq. [1])
Hiller, D.M., and Crabb, G.I. 1998. Prediction of groundborne vi- k1, k2, k3 constants of proportionality (eq. [2])
bration from mechanised construction works. Proceedings of the l depth of pile penetration
Institute of Acoustics, 20(2): 25–32. n attenuation index (eq. [7])
Hiller, D.M., and Hope, V.S. 1998. Groundborne vibration gener- r horizontal radial offset distance from pile shaft
ated by mechanised construction activities. Proceedings of the r0 reference offset distance
Institution of Civil Engineers, Geotechnical Engineering, s slope distance from pile toe
131(4): 223–232. T ratio of energy transmission to reflectance at a boundary
Hosking, I.A., Mirkov, P., and Deb, P.K. 1988. Predicted, measured for normal incidence
ground vibrations caused by pile driving in sand at the Tuncurry v0 peak particle velocity at reference offset r0
Unloading Wharf, New South Wales. In Proceedings of the 5th vv peak vertical particle velocity
Australia – New Zealand Conference on Geomechanics, Sydney, vr peak particle velocity at offset r
Australia. pp. 421–426. vres peak resultant particle velocity
Jongmans, D. 1996. Prediction of ground vibrations caused by pile
W input energy term (eqs. [1] and [2])
driving: a new methodology. Engineering Geology, 42: 25–36.
Web theoretical vibrational energy from transmission across
Lo, M.B. 1977. Attenuation of ground vibration induced by pile
pile toe – soil boundary
driving. In Proceedings of the 9th International Conference on
Soil Mechanics and Foundation Engineering, Tokyo, Vol. 4, Wf theoretical vibrational energy due to shaft friction ef-
pp. 281–286. fects
Luna, W.A. 1967. Ground vibrations due to pile driving. Founda- WH impact hammer energy per blow
tion Facts, Vol. 3, No. 2, Raymond International, Houston, Tex. x index (eq. [1])
Poulos, H.G., and Davis, E.H. 1980. Pile foundation analysis and Z acoustic impedance of a material
design. John Wiley and Sons, New York. α attenuation coefficient, associated with material damping
Telford, W.M., Geldart, L.P., Sheriff, R.E., and Keys, D.A. 1976. ρ mass density
Applied geophysics. Cambridge University Press, Cambridge,
U.K.

© 2000 NRC Canada

I:\cgj\Cgj37\Cgj03\T99-130.vp
Wednesday, June 28, 2000 3:31:40 PM

You might also like