1 s2.0 S0267726122003554 Main - Compressed

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering


journal homepage: www.elsevier.com/locate/soildyn

Vibration impact of rock excavation on nearby sensitive buildings: An


assessment framework
Shiguang Wang a, Songye Zhu b, *
a
Institute for Interdisciplinary and Innovate Research, The Xi’an University of Architecture and Technology, Xi’an, China
b
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Hong Kong, China

A R T I C L E I N F O A B S T R A C T

Keywords: As a common construction activity, rock excavation using hydraulic breakers often generates high-level ground-
Rock excavation and structure-borne vibrations, which may adversely affect nearby buildings in terms of structural safety,
Building vibrations occupant comfort, and ultraprecision equipment functionality. However, building vibrations induced by rock
Mechanical impedance
excavation have not been well investigated and understood in the literature. For example, the coupling atten­
Impedance-based prediction model
Field measurement
uation at column bases and the floor-to-floor vibration attenuation inside a building have rarely been quanti­
Vibration impact assessment tatively studied. Moreover, traditional dynamic analysis methods in earthquake engineering may not apply to
simulations of rock excavation-induced building vibrations. This paper proposed a novel assessment framework
based on the mechanical impedance concept, which can quantify the coupling attenuation at column bases and
the floor-to-floor vibration attenuation inside a building in a computationally efficient manner. Systematic finite
element simulations of rock excavation were performed to characterize vibration propagation in the ground and
buildings. The finite element simulations and field measurements successfully verified the accuracy of the
proposed prediction formulas, which offers a convenient vibration impact assessment framework to construction
practitioners.

1. Introduction et al. [5] demonstrated that blasting-induced vibrations in a 22-story


high-rise building attenuated on the upper floors compared with input
In densely populated metropolises, construction activities (such as base motions. Such floor-to-floor attenuations were also widely
piling and rock excavation) often produce negative vibration impacts on observed in train-induced building vibrations. For example, Xia et al. [6]
nearby building structures as well as on occupants and ultraprecision and Zou et al. [7] observed steady, monotonic reduction in velocity
equipment inside these buildings. Field measurements of vibration levels with increasing floor elevations. Such attenuations are mainly due
levels on the ground surface and inside buildings are commonly per­ to the dispersion of vibration energy in its upward propagation [8].
formed during construction processes. Ground-borne vibrations atten­ In addition to ground-level (outdoor) and floor-to-floor (indoor) at­
uate with plan distances due to geometric and material damping. tenuations, vibration attenuation occurs at the foundation level. Sanayei
Vibration propagations along the free ground surface have been exten­ et al. [9] measured train-induced vibrations and investigated how vi­
sively studied [1]. bration attenuated from the open fields adjacent to the building to the
Inside buildings, vertical floor vibrations are typically much larger building foundations. This effect is sometimes referred to as coupling
than lateral vibrations [1,2]. The dynamic properties of buildings and attenuation [8], which implies that such attenuation is due to
the characteristics of excitation sources jointly determine building vi­ ground–foundation coupling. Notably, it is different from normal
brations. If excitation frequency is close to the natural frequencies of soil–structure interactions as investigated in previous studies (e.g.,
building floors, vibration amplifications on higher floors of multistory Auersch [10] Sanitate and Talbot [11]) because foundations suppress
buildings may occur compared with vibration intensity measured on the ground vibrations in the coupling attenuation. The coupling attenuation
ground surface [3]. However, impulsive excitations or excitations with a depends highly on vibration frequency and can often be ignored for
limited number of cycles generated by construction activities can hardly low-frequency excitations. For example, it is always assumed in earth­
excite highly amplified resonant responses [4]. For example, Oriard quake engineering that ground motions are independent with

* Corresponding author.
E-mail address: songye.zhu@polyu.edu.hk (S. Zhu).

https://doi.org/10.1016/j.soildyn.2022.107508
Received 5 April 2022; Received in revised form 24 July 2022; Accepted 20 August 2022
Available online 1 September 2022
0267-7261/© 2022 Elsevier Ltd. All rights reserved.
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

superstructures [12,13], which essentially ignores the coupling attenu­ content than other activities, are still not well investigated and fully
ation effect in the analyses. understood. (2) Traditional dynamic analysis methods in earthquake
Various guidelines (e.g., Caltrans [1]) present empirical formulas engineering cannot be directly adopted to predict building floor vibra­
derived from engineering tests and numerical simulations to predict tions induced by rock excavation. (3) A sophisticated FE model of vi­
vibrations on free ground surfaces induced by various construction ac­ bration sources, ground, foundations, and buildings can systematically
tivities. However, they usually do not quantify the aforementioned and accurately characterize vibration propagations and attenuations in
coupling and floor-to-floor attenuations. An informative guideline is different stages. However, construction practitioners cannot adopt such
from the Federal Transit Administration (FTA) [8] for assessing the noise a computationally expensive approach easily and quickly.
and vibration impacts induced by transit projects, which considers many In summary, a simple assessment method, which can be used by
factors, including propagation distances, building foundations, floor engineers to predict vibrations induced by rock excavation quickly, re­
numbers, and material damping, as well as speeds, wheel types, rail, and mains absent so far. To fill this gap, this paper proposed a quantitative
track support system. However, adjustment factors in FTA [8] for these assessment framework to predict building vibrations induced by short-
factors are defined as single numbers without considering the different duration impact generated by rock excavation work using hydraulic
dynamic characteristics of foundations and buildings. Such an over­ breakers. First, complete vibration transmission paths (including
simplified approach cannot predict vibration very accurately. ground- and structure-borne vibrations) were characterized based on the
Numerical modeling and simulations are also commonly used to impedance concept. Impedance-based theoretical formulas were derived
predict building vibrations induced by construction activities. For to predict the vibration coupling attenuation at column bases and the
example, Zhu et al. [14] built the finite element (FE) model of a building floor-to-floor vibration attenuation inside a multistory building. Subse­
and simulated floor vibrations induced by piling activities, where scaled quently, theoretical definitions of point driving impedance for ground
ground motions were input to column bases. This simulation method has and structural components were introduced. To verify the efficacy of the
its root in earthquake engineering. However, this traditional method proposed formulas, transient vibrations propagating into column bases
becomes inaccurate and problematic when applied to simulations of and upper floors were systematically investigated through the FE sim­
construction-induced vibrations: (1) This method assumes that input ulations of impact waves due to rock excavation work. Moreover, the
ground motions are determined by excitation sources and independent proposed formulas were verified by comparing them with field mea­
of building structures. However, the existence of foundations clearly surement data. The predicted vibration levels can be further compared
suppresses underneath ground vibrations compared with those on free with tolerable vibration limits to assess the potential influence on
ground surfaces, as discussed in coupling attenuation. (2) Ground vi­ structures, occupants, and ultraprecision equipment inside buildings.
brations attenuate rapidly with plan distances due to material and
geometric damping. Therefore, all column bases are subjected to ground 2. Vibration transmission paths
vibration inputs with different magnitudes, phase shifts, and directions,
which presents the difficulty in simulations. Fig. 1 shows the vibration transmissions from the ground to a su­
More accurate numerical simulations can be performed by simu­ perstructure. Vibrations generated by a source first propagate horizon­
lating vibration generation, propagation, and reception. Chua et al. [15] tally along the ground surface to a plan distance d and then transmit
simulated the subway–soil–structure system to predict ground-borne upward along the columns in a multistory building. As construction-
vibrations in a 4-story podium block due to subway trains. Lopes et al. induced vibrations are typically dominated by vertical components,
[16] developed a comprehensive model including train, tunnel–ground, this paper will mainly focus on the vertical vibration velocities of
and buildings to predict underground railway-induced vibrations in an structural members.
urban environment. Yang et al. [17] modeled the The transmission path has two wave division mechanisms. One is the
track–tunnel–soil–building system and proposed a two-step time­ coupling attenuation at a column base, which leads to the column base
–frequency method to predict train-induced vibrations. However, this velocity Vbase smaller than the velocity Vref at a reference point on a free
detailed modeling approach is computationally expensive, and its ground surface, where the column base and the reference point have an
case-specific results are not applicable to vibration assessment in generic equal distance d from the vibration source. The other is floor-to-floor
cases. attenuation at column–beam–slab connections because part of vibra­
The impedance-based approach arising from wave theory offers a tion energy is dispersed into beams and slabs in the upward transmission
more efficient alternative for predicting building vibration levels, in procedure. The impedances of different components are used in this
which numerous simplifying assumptions are often made to simplify the paper to quantify these two attenuation mechanisms and estimate
computation of mechanical impedance. For example, Auersch [18]
developed a simple building model and predicted building–soil reso­
nances by attaching a building mass to the dynamic stiffness matrix of
layered soils. Sanayei et al. [19] developed a simple one-dimensional
(1D) impedance-based analytical model to calculate train-induced vi­
brations. Consequently, each finite segment of a column was charac­
terized by the impedances at the top and bottom of the segment, while
each floor’s impedance was the input impedance at the interface be­
tween the floor and the column. Zou et al. [20] added load-bearing walls
with a finite width into the above 1D impedance model, in which the
wall segments were treated as a column with the same cross-sectional
area. A 2D impedance model was developed to consider both bending
and axial waves transmitting from foundations to upper floors in
train-induced vibrations, thus describing vibration transmission in col­
umns, load-bearing walls, and floors.
Several main limitations can be identified in the above literature: (1)
Many past studies were focused on continuous or transient vibrations
induced by railway traffics, piling, or blasting. Vibrations induced by
rock excavation using hydraulic breakers, which have an extremely
short duration (around 0.5 ms only) and can generate higher frequency Fig. 1. Vibration transmissions from the ground to a superstructure.

2
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

structural vibrations due to a nearby vibration source, where mechani­ vibration decreases with the increasing separation distance because
cal impedance is defined as the ratio of the force to the velocity. The high-frequency contents are damped more rapidly than low-frequency
impedances of the ground, column, beam, and slab are denoted by zg, zc, contents.
zb, and zs, respectively, in Fig. 1.

2.1. Ground vibration propagation 2.2. Coupling attenuation at column bases

Empirical formulas are commonly used to predict vibration attenu­ When the column base is excited by an incident ground wave, the
ation along a free ground surface due to geometric and material driving force Fc at the column base equals column base velocity Vbase
damping. For example, Wang et al. [21] investigated the ground-borne multiplied by column impedance zc.
vibrations induced by rock excavation using percussion methods, Fc = Vbase zc , (5)
wherein the ground surface vibration was quantified as peak particle
velocity (PPV) considering different ground damping ratios and/or rock Impedance zc corresponds to P waves in the column because the
types. Accordingly, Vref can be determined by the following empirical vertical vibrations are of interest in this paper. Meanwhile, the column
formulas: base also applies reaction force Fc on the ground, which generates
√̅̅̅̅̅̅̅̅̅̅̅̅ another velocity component on the ground and makes the ground vi­
Vref = kr × ηErated × (d + ar )− nr bration at the column base different from that on a free ground surface.
(1)
Lref (VdB) = 20 × log 10 Vref + 92
Fc = V R z g , (6)
where Vref (unit: mm/s) is the ground velocity induced by a hydraulic
breaker with the rated energy Erated (unit: J), Lref is the logarithm of Vref Vbase = Vref − VR , (7)
(in VdB re: 1 μ-in/s), η = 33%–50% is the energy transfer efficiency that
is defined as the ratio of the actual input energy to the rated energy, where VR is a new velocity component generated by Fc on the ground,
d (unit: m) is the distance between the vibration source and the point of and Vref is the velocity generated by the vibration source on a free
interest, and three other parameters: intensity coefficient kr (unit: mm/ ground surface (without column).
s/m− n/J0.5), attenuation factor nr, and distance constant ar (unit: m) Based on Equations (5)–(7), a complex velocity ratio Vbase/Vref can be
depending on ground material damping ratio and/or rock types. More derived as a function of the impedance ratio of zc/zg.
details about these parameters were reported by Wang et al. [21]. Vbase 1
In addition to vibration amplitude attenuation, the frequency con­ = , (8)
Vref 1 + zzgc
tent of ground-borne vibrations varies during propagation, which was
not discussed by Wang et al. [21]. This paper used the following formula where ground impedance zg is a complex number; thus, the velocity ratio
to approximate the frequency content of an impact source: is also complex. Assuming ground impedance with an amplitude |zg| and
(
π
) a phase angle φ1, then the coupling attenuation ratio VRb between the
|y| = y0 sin ω , 0 ≤ ω ≤ 2ω0 , (2) column base velocity and the reference point velocity can be expressed
2ω0
as

⎛ ⎞
⃒ ⃒ ⎜ ⎟
⃒ 1 ⃒ ⎜ 1 ⎟
VRb (VdB) = 20 × log 10 ⃒⃒ ⃒ = 20 × log 10 ⎜√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
̅ ⎟, (9)
1 + zzgc ⃒ ⎜
⎝ 1 + zc
( )2 ⎟
+ 2 z cos(180 − φ1 ) ⎠
zc
|zg | | g|

where y0 refers to the velocity amplitude at the dominant frequency


ω0 of the impact source, and y is the amplitude at a frequency ω. Based where the attenuation ratio is expressed in decibels (VdB). In general, a
on the field measurement described in Section 5, ω0 = 9425 rad/s (i.e., larger impedance ratio (zc/|zg|) leads to a larger attenuation due to
f0 = 1500 Hz) is adopted in this paper for the impact load induced by a coupling loss.
hydraulic breaker.
Considering a shear wave velocity cs and material damping ratio ξ of
2.3. Floor-to-floor attenuation
the ground, the material damping effect is frequency dependent in vi­
bration propagation [22], whereas geometric damping is constant for a
Wave propagation from one medium with impedance Z1 to another
specific wave type. Consequently, the frequency content of ground vi­
medium with impedance Z2 was well investigated in the literature [23].
bration can be expressed as
In a special 1D propagation case with a normal incident angle, the
[ ( )]
π transmission coefficient T is defined as
(3)
ωξd
|Y| = y0 sin ω × d− 0.5 × e− cs ,
2ω0
A2 2Z1
T= = , (10)
where |Y| is the amplitude of frequency ω at a location with distance A0 Z1 + Z2
d from the impact source. By searching the maximum |Y|, the dominant
where A0 is the vibration amplitude of the incident wave, and A2 is the
frequency at a distance d from an impact source can be expressed as
vibration amplitude of the transmitted wave.
( )
2ω π cs At the column–beam–slab connection of each floor in a building
ωeq = 0 arctan , (4)
π 2ω0 ξd (Fig. 1), wave propagation can be simplified as a transmission from the
lower column segment with impedance zc to the upper colum­
Equation (4) indicates that the dominant frequency of the ground n–beam–slab connection with increased impedance zc + z. This

3
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

simplification neglects the actual connection dimensions and multiple in the Appendix.
reflection effects between two nearby connections. Consequently, a
floor-to-floor velocity amplitude ratio VRf2f-1 (VdB) due to energy divi­ 3. Definitions of point driving impedance
sion at the column–beam–slab connection for each floor can be obtained
by substituting Z1 = zc and Z2 = zc + z into Equation (10). This section presents the impedance of the ground and structural
⃒ ⃒ members to be used in Equations (9) and (11). Although impedance was
⃒ 1 ⃒ 1
VRf2f− 1 (VdB)=20×log 10 ⃒⃒ ⃒
z ⃒ = 20×log 10
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ̅, derived in some ideal situations considering infinite structure di­
1+ 2zc |z|
1+ 2zc +2 2zc cosφ2 |z| mensions and harmonic responses, the current paper shows that it can be
adopted to describe the attenuation of impact wave-induced transient
(11) vibrations inside a general building with reasonable modifications.
where z is the sum of the impedance contributions from the floor slabs
3.1. Ground impedance
and beams. Assume zs is the impedance of a floor slab in one quadrant,
and zb is the impedance of a beam on one side. The value of z depends on
Assume a rectangular foundation of area Ac = b × c on a half-space
the numbers of beams and slabs connected to the column: z = 4zs + 4zb
surface is subjected to a uniform dynamic load, as shown in Fig. 2.
for an internal column, z = 2zs + 3zb for a boundary column, and z = zs +
Thomson and Kobori [24] analytically derived the vertical nondimen­
2zb for a corner column. As shown in the following section, slab
sional compliance at the center of the rectangular foundation as a
impedance zs is a real number, and beam impedance zb is a complex
function of the nondimensional frequency. By adopting the same deri­
number. Thus, φ2 is the phase angle of total impedance z. φ2 = 0 when
vation, this paper further defines the nondimensional ground impedance
only slab impedance zs is considered, and φ2 = 45◦ when only beam
of zg/(Acρcs) based on the impedance definition.
impedance zb is considered.
⎡ ⎡ ( )⎤ ⎤
In addition to wave energy division at column–beam–slab connec­ / ) (
∫ ∞ ∫ π2 2 ωb ωc
tions, material damping effects also contribute to the floor-to-floor
1
zg ⎢ 4j (z − n ) ⎢
2 2 sin 2cs
z cos θ sin 2cs
z sin θ ⎥ ⎥
=1 ⎢ ⎢ ⎥dθdz⎥,
attenuation in upward vibration transmission. Considering a compres­ Ac ρc s ⎣π 2
0 0 F(z) ⎣ z sin θ cos θ ⎦ ⎦
sive wave in the column, the damping effect is a function of wave ve­
locity, frequency, damping ratio, and propagation distance [22]. The (15a)
corresponding attenuation between two adjacent floors (i.e., from the
( )2 ( )1 ( )1
ith floor to (i+1)th floor) caused by material damping can be calculated F(z) = 2z2 − 1 − 4z2 z2 − n2 2 z2 − 1 2 , (15b)
by
( ) √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
Vi+1 ( − ωeq ξc h0 ) cs 1 − 2υ
VRf2f− 2 (VdB) = 20 × log 10 = 20 × log 10 e cp n= = , (15c)
Vi cd 2(1 − υ)
ωeq ξc h0
= − 8.68 × , (12)
cp where ρ, cs, and ω refer to the ground density, ground shear wave ve­
locity, and excitation circular frequency, respectively; n is the ratio of
where cp is the compressive wave velocity in the concrete column, ξc is shear wave velocity cs to dilatation wave velocity cd, and it is related to
the material damping ratio in the column, and h0 is the floor height. the Poisson’s ratio υ of the ground.
The total floor-to-floor attenuation is the sum of the above two ef­ Owing to the existence of Rayleigh poles z0 where the denominator
fects. of the integrand equals zero [i.e., F(z0) = 0], the analytic solution of
Equation (11) is difficult, but numerical evaluations can be performed
VRf2f = VRf2f− 1 + VRf2f− (13)
2
[24]. After numerical evaluations of Equation (15), Fig. 3 shows the
nondimensional ground impedance (including amplitude |zg|/(Acρcs)
and phase angle φ1) versus nondimensional frequency (ωA0.5 c /cs) for a
2.4. Building vibration prediction and impact assessment
square foundation (b = c).
Considering the numerical solutions of Equation (15) remain too
Considering ground propagation, coupling attenuation, and floor-to-
complicated to be adopted in engineering applications, a simplified
floor attenuation, the vibrations of the ith floor of a building induced by
rock excavation using a hydraulic breaker can be expressed as
Lv = Lref + VRb + i × VRf2f , (14)

where Lv (VdB) is the vibration velocity of the target floor, Lref (VdB) is
the ground vibration velocity at a reference point on the free ground
surface, and i refers to the floor number and starts from i = 0 for the
ground floor. In the construction industry, Lv and Lref are typically
expressed as the logarithm of PPV.
Subsequently, vibration impact assessment can be conducted by
comparing the predicted vibrations with the tolerable vibration limits
considering the occupants and equipment inside a building. Notably, the
tolerable vibration limits are typically expressed as root–mean–square
(RMS) velocity, instead of PPV used by the construction industry. In
practice, a crest factor (CF = PPV/Vrms) is commonly adopted to esti­
mate the relationship between two indices [1]. Consequently, PPV (VdB)
can be converted to Vrms (VdB) by deducing Lv by 20 × log10 CF. FTA [8]
suggested the Vrms limits (in VdB re: 1 μ-in/s) for three categories of
buildings, namely, 65 dB for the operation of sensitive equipment, 72 dB
for normally sleeping people, and 75 dB for primarily daytime use. The
complete vibration prediction and assessment framework is summarized Fig. 2. Uniform load area on the half-space.

4
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Fig. 3. Nondimensional relation of (a) ground impedance amplitude |zg|/(Acρcs), and (b) phase angle φ1 versus nondimensional frequency ωA0.5
c /cs for square
foundation (b = c).

equation of zg is proposed based on regression analysis. Axial vibrations in the column lead to bending vibrations in the floor
⃒ ⃒ ⃒ ⃒ slab at connections. The axial load in the column and the floor slab can
zg = R + jI = ⃒zg ⃒cos φ1 + j⃒zg ⃒sin φ1 , (16a)
be approximately represented by a point load on a thin infinite plate
[26]. According to Kirchhoff plate theory [27], driving point impedance
R = − 0.81Ac ρcs , (16b)
zs, which is the ratio of the point load to the velocity at the driving point,
can be calculated as
2.06ρA0.5
c cs
2
I= , (16c) √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
ω Eρ
̂z s = 8s 2
( s s ), (18)
⃒ ⃒ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )− 2
̅ 12 1 − υ2s
⃒zg ⃒ 2.062 cs 2 ωAc 0.5
= 0.81 +2 = 0.656 + 4.244 , (16d)
Ac ρcs ω 2 Ac cs where s is the slab thickness; ρs, Es, and υs are the slab density, Young’s
modulus, and Poisson’s ratio, respectively. For a slab in one quadrant,
0.393ωAc 0.5 the slab impedance is only one-quarter of Equation (18). Considering
φ1 = arctan + 90 (16e)
cs that the real slab conditions deviate from the ideal assumptions, the
impedance of a slab in one quadrant can be approximately expressed as
where the unit of the phase angle φ1 is degree, |zg| is the magnitude of √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
the ground impedance, R represents the damping coefficient that is Eρ
zs = αs ⋅2s2 ( s s ), (19)
frequency independent, and I represents the stiffness effect that is fre­ 12 1 − υ2s
quency dependent and tends to be zero at a high frequency. This
observation is similar to that for a rigid foundation with a circular sec­ where the modification coefficient αs will be determined later based on
tion [25]. Fig. 3 indicates that the regression equations agree with the FE simulations.
accurate numerical solutions for a square area. The variations of ground Similarly, a column–beam connection can be approximated by an
impedance (|zg| and φ1) with various parameters (excitation frequency isotropic, infinite beam excited by a transverse point load. The mobility
ω, shear wave velocity cs, and foundation area Ac) were also investigated function Yb evaluated at the driving point was defined by Fahy and
through a series of FE simulations. The simulation results also verify the Gardonio [28]. Accordingly, the driving point impedance zb of an
accuracy of Equation (16) proposed in this paper. The corresponding FE infinite beam can be derived as the reciprocal of mobility function Yb as
results are not shown here due to the page limitation. For a rectangular
area in Fig. 2, the accurate solution has to be computed numerically, but 2
(20)
3 1 3 1
̂z b = bb hb 2 Eb 4 ρb 4 ω2 (1 + j),
it can be approximated by a square section with an equal area Ac with 121/4
acceptable accuracy. where ω is the circular driving frequency; ρb and Eb are the beam density
Equation (16d) indicates that the ground impedance zg becomes very and Young’s modulus, respectively; bb and hb are the horizontal and
large at a low frequency, and consequently, the coupling attenuation vertical dimensions of the beam cross-section, respectively. Considering
ratio VRb in Equation (9) becomes zero. This explains why such coupling a beam only on one side of the column and the deviation of an actual
attenuation is commonly ignored in earthquake engineering, but it has beam from an idealized beam, the impedance of a beam can be
to be considered in rock excavation induced vibrations given much approximately expressed as
higher frequency contents.
αb
(21)
3 1 3 1
zb = bb hb 2 Eb 4 ρb 4 ω2 (1 + j),
121/4
3.2. Structural member impedance
where the modification coefficient αb will be determined later based on
The driving point axial wave impedance of an infinite column (i.e., FE simulations.
zc) is expressed as [26]. Notably, the impedance of the column and slab (i.e., zc and zs) are
√̅̅̅̅̅̅̅̅̅ real numbers, and the impedance of the ground and beam (i.e., zg and zb)
z c = Ac Ec ρ c , (17) are frequency-dependent complex numbers. In addition, the floor-to-
floor attenuation VRf2f-2 also depends on vibration frequency. It jus­
where Ec, ρc, and Ac are the Young’s modulus, mass density, and cross- tifies the need for an empirical formula, that is Equation (4), to estimate
sectional area of the column, respectively.

5
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

the equivalent frequency in vibration transmission before performing Table 1


impedance-based calculations. Ground material properties.
Parameters Six sets of rock mass (R1–R6)
4. Numerical verification via FE modeling
R1 R2 R3 R4 R5 R6

4.1. Mesh and geometry Young’s modulus: E (GPa) 1.4 9 20 30 42 60


Shear wave velocity: cs (m/s) 473 1200 1790 2192 2594 3100
Density: ρ (kg/m3) 2600
Fig. 4 shows the FE modeling of a full-scale 12-story frame building Poisson’s ratio: υ 0.2
on a semi-infinite ground domain established in commercial software
ABAQUS. The simulated ground domain was 90 m long (X) × 70 m wide spacing was 6 m in the X and Y directions, and the height of each story
(Y) × 60 m deep (Z). The ground domain was modeled using the element was 4 m.
type C3D8R. Fine mesh sizes were adopted for the interior ground
domain, whereas the exterior ground domain was discretized into a
graded mesh to reduce computation cost. In the interior domain, 4.2. Material and section parameters
different mesh sizes (80, 200, 250, 300, 400, 500 mm) were used for six
studied rock mass sets (i.e., R1–R6 in Table 1). In particular, finer Hoek and Brown [29] defined three sets of rock mass (namely, very
meshing was defined in the area near the impact location. Surrounding poor quality rock mass, average quality rock mass, and very good quality
four sides of the ground domain were 10-m quiet boundaries modeled hard rock mass) based on the moduli of elasticity, which correspond to
and discretized into infinite elements (CIN3D8). the rock sets R1, R2, and R5, respectively, in this paper. Table 1 lists that
In the top view of the model, the dotted rectangular box 42 m long three additional rock sets, namely, R3, R4, and R6, were considered to
(X) × 12 m wide (Y) shows the location of the 12-story frame building investigate the effects of different ground conditions better. Table 1
consisting of columns, beams, and slabs. The element type C3D8R with a presents the corresponding Young’s moduli, shear wave velocities,
size of 200 mm was used to model the frame building. The column densities, and Poisson’s ratios.
Concrete is modeled as an elastic material, considering the building

Fig. 4. FE model of a 12-story frame structure on the ground domain: (a) isometric view of the full model, (b) enlarged view showing the impact location, and (c) top
view of the full model.

6
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Table 2
Building structural components.
Structural Cross-section Values
components dimensions

Column b (m) × c (m) 0.25 × 0.25, 0.5 × 0.5, 0.75 × 0.75, 1.0 ×
1.0, 1.5 × 1.5, 2.0 × 2.0
Beam bb (m) × hb (m) 0 × 0 (no beam), 0.3 × 0.2, 0.3 × 0.4, 0.3
× 0.6, 0.3 × 0.8, 0.2 × 0.4, 0.4 × 0.4
Slab thickness s (m) 0 (no slab), 0.075, 0.15, 0.20, 0.25

vibrates at a relatively low strain level. The density and Young’s


modulus of concrete components were defined in the normal range of
2000–2800 kg/m3 and 10–50 GPa, respectively. A constant Poisson’s
ratio of 0.2 was adopted. Rayleigh damping was adopted in the
modeling to consider a damping ratio of 5% for the ground and 3% for
the superstructure.
Table 2 lists different dimensions of structural components simulated
in this paper, including six column cross-sections, eight beam cross-
sections, and five slab thicknesses, covering a general range of each
structural component. The actual sizes of the column, beam, and slab of
a reinforced concrete structure depend on structural design considering
all types of loading on it.

4.3. Dynamic loading and simulation

The authors conducted field measurements to quantify the impact


strain on the steel chisel tool when a hydraulic breaker strikes concrete
ground.
Fig. 5(a) shows the measured impact strain of a steel chisel when a
hydraulic breaker excavates the ground. The measurement details are
presented in Section 5. The time history contained around 10 impact
shocks in 1 s. Fig. 5(b) shows the average striking force of the chisel in
one shock, which was calculated based on the measured strain, Young’s
modulus, and cross-section area. The duration of an impact impulse was
less than 700 μs, considerably shorter than the long interval (around
100 ms) between two consecutive impacts. Therefore, the superposition
of the vibrations induced by two consecutive impacts is unlikely to
occur.
The impact force was applied as a uniform pressure on an area with a
40 mm diameter of the chisel on the ground surface. In addition to the
impact force, harmonic excitations were simulated for comparison.
Transient analyses using an explicit time integration method were per­ Fig. 6. Steady-state amplitude spectra at (a) column-beam connections, and (b)
formed to investigate the transient responses of the model induced by slab centers.
rock excavation impacts. The incremental time step was automatically
selected by ABAQUS based on the minimum mesh size in the model. 4.4. Transient vibration in the ground and superstructure

Before the simulations of transient vibrations, the modal dynamic


analysis and steady-state analysis of the established building were

(a) (b)
Fig. 5. Measured impact load generated by a hydraulic breaker: (a) impact strains and (b) average impact force.
7
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Fig. 7. Simulated field of vibration intensity of velocity (mm/s) showing (a) wave propagation from ground to building, and (b) wave propagation inside
the building.

performed. The modal analysis indicated that the fundamental vibration simulated. Equation (9) indicates that VRb is a function of multiple
modes of the columns and floor slabs correspond to the frequencies of factors. Fig. 9(a) and (b) show the variations of VRb with the ground
5.9 Hz and 12.1 Hz, respectively. Considering a free-field harmonic shear wave velocity cs and column area Ac, respectively. The harmonic
incident wave with a unit magnitude at one column base, the steady- excitation simulations match the theoretical predictions accurately,
state responses represent the ground-to-structure transfer function. whereas the impact excitation cases slightly deviate from the theoretical
Fig. 6(a) and (b) show the transfer functions at a column-beam predictions. However, in general, the theoretical predictions using the
connection and at the center of a floor slab, respectively. Resonance equivalent frequency feq of the impact simulations still have satisfactory
induced vibration amplifications could be observed only at two funda­ accuracy. The effects of other parameters in Equation (9) were also
mental vibration modes (around 6 Hz at both column or floor points, and verified through a series of FE simulations but are not presented due to
around 12 Hz at floor center). In general, with the increasing frequency the page limit.
and floor height, both transfer functions showed strong attenuation.
These phenomena are consistent with the previous study (e.g., Ref. [30]) 4.6. Verification of floor-to-floor attenuation
that reported that the resonance of structural members only occurred at
low frequencies. Fig. 10 shows the example of simulated velocity time histories on
Fig. 7 shows the transient simulations of vibration propagation in different floors, where the responses at the column points (connections)
terms of the velocity magnitude in the ground and building. Vibration and the centers of the adjacent floor slabs are presented. The PPV value
propagations in the ground domain consist of body waves below the on each floor is always induced by the incident wave instead of the re­
ground surface and surface waves near the ground surface, where the flected wave. The PPV can be easily identified in the simulated time
latter exhibits a larger amplitude. Parts of the impact wave propagate windows. The corresponding amplitude spectra at column points and
from the ground to the column bases and then upward along the col­ slab centers are shown in Fig. 11(a) and Fig. 11(b), respectively. Com­
umns with wave divisions at the column–beam–slab connections. parisons of vibration amplitudes in both time and frequency domains
show that the vibration amplitude at slab centers was smaller than the
4.5. Verification of coupling attenuation at column bases corresponding column points on different floors. The simulated vibra­
tions mainly show amplitude attenuation at high-frequency along the
Fig. 8 shows the vertical velocity time histories and the corre­ building height and the horizontal plane at each floor (from column
sponding amplitude spectra of two selected column bases with distances points to slab centers).
d = 2 m and d = 10 m from the impact source. Compared with the vi­ The resonant frequencies (6 Hz and 12 Hz) identified from the
bration responses at the reference points on a free ground surface with steady-state analysis in Fig. 6 could not be observed in Fig. 11. The
the same distances, apparent attenuation of velocity intensity at the impulsive excitations with an extremely short duration and high-
column bases can be observed due to the coupling effect. Impact load frequency contents could hardly excite amplified resonant responses at
excites vibrations in a wide frequency bandwidth. Higher frequency low frequencies.
attenuates faster in ground propagation due to the material damping The floor-to-floor attenuation of PPV in Equation (13) consists of two
effect. As predicted by Equation (4), the dominant frequency decreases parts, VRf2f-1 and VRf2f-2. This subsection mainly focuses on VRf2f-1 at
from 1240 Hz at d = 2 m to 600 Hz at d = 10 m. column–beam–slab connections because VRf2f-2 due to material damp­
Coupling attenuation VRb can be evaluated by comparing the ve­ ing is well understood. Three structural configurations, namely, column
locities at the column bases and corresponding reference points. Fig. 9 + slabs, columns + beams, and columns + beams + slabs, were simu­
compares the simulation results with the theoretical prediction by lated to quantify the floor-to-floor attenuation. The simulated average
Equation (9), in which both the impact and harmonic excitations were VRf2f values of 10 floors of the building were compared with the

8
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Fig. 10. Transient velocity time histories at column points (connections) and
slab centers on different floors (column at d = 10 m).

predictions by Equations (11) and (12).


Considering an interior column in the first beam–slab configuration
(i.e., the impedance z = 4zs), VRf2f-1 becomes a function of the slab-to-
column impedance ratio zs/zc. Based on a baseline case (Ac = 0.5 m ×
0.5 m, s = 0.15 m, feq = 600 Hz, Ec = Es = 30 GPa, and ρc = ρs = 2400 kg/
m3), parametric studies were conducted in the FE simulations by varying
individual parameters: slab thickness s (0–0.25 m), equivalent frequency
feq (50–1200 Hz), Young’s modulus of concrete Ec (10–50 GPa), and
Fig. 8. Vertical velocity time histories and corresponding amplitude spectra for concrete density ρc (2000–2800 kg/m3). Fig. 12 shows that the simu­
column bases at two distances: (a) d = 2 m and (b) d = 10 m (Simulation case: lated VRf2f-1 is strongly correlated with the impedance ratio zs/zc. The
rock set R4 with cs = 2192 m/s and column with b × c = 0.5 m × 0.5 m, Ec = 30 change in the slab thickness s considerably influences slab impedance
GPa, ρc = 2400 kg/m3). ratio [zs ∝ s2 according to Equation (19)], whereas variations in equiv­
alent frequency feq, and concrete properties Ec and ρc have minimal

(a) (b)
Fig. 9. Comparison between the theoretical predictions and numerical simulations of coupling attenuation VRb for (a) different ground shear wave velocity cs (f =
400 Hz, Ac = 0.25 m2), and (b) different foundation area Ac (m2) (f = 400 Hz, cs = 2192 m/s).

9
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Fig. 13. Variations of VRf2f-1 (VdB) with the impedance ratio of the beam-to-
column |zb|/zc.

effect on the impedance ratio. Fig. 12 illustrates the theoretical pre­


dictions of VRf2f-1 by Equation (11) using αs = 1.0 and αs = 1.3. The
curve with αs = 1.0 underestimates the attenuation apparently, whereas
the curve with αs = 1.3 agrees with all the simulated cases very well,
which justifies the need for a modification coefficient αs in Equation
(19).
In the second configuration with only columns and beams, the
attenuation ratio VRf2f-1 is controlled by the beam-to-column impedance
ratio |zb|/zc. Considering an interior column connected to four beams,
the beam impedance is z = 4zb. Based on a baseline case (Ac = 0.5 m ×
0.5 m, bb = 0.3 m, hb = 0.4 m, feq = 600 Hz, Ec = Eb = 30 GPa, and ρc = ρb
= 2400 kg/m3), parametric studies were conducted in the FE simula­
tions by varying individual parameters: beam cross-section dimensions
bb (0–0.4 m) and hb (0–0.8 m), equivalent frequency feq (50–1200 Hz),
concrete’s Young’s modulus Ec (10–50 GPa), and concrete’s density ρc
(2000–2800 kg/m3). Fig. 13 shows the simulated VRf2f-1 is strongly
correlated with the impedance ratio |zb|/zc. According to Equation (21),
Fig. 11. Transient amplitude spectra at (a) column points (connections), and zb ∝ b, zb ∝ h2/3, and zb ∝ f1/2
eq . Therefore, the changes in bb, hb, and feq
(b) slab centers. considerably affect the impedance ratio. By contrast, the variations of
concrete properties Ec and ρc only slightly modify the impedance ratio.
Fig. 13 compares the theoretical predictions of VRf2f-1 by Equation (11)
with the FE simulations, where the theoretic curves correspond to αb =

Fig. 12. Variations of VRf2f-1 (VdB) with the impedance ratio of slab to column Fig. 14. Comparisons of the simulated VRf2f (VdB) values with the theoretical
zs/zc. predictions in different cases.

10
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Fig. 15. Photos of field measurements: (a) one vibration measurement point on the ground surface, (b) one vibration measurement point on a column base, (c) five
vibration measurement points on a corner column, and (d) a strain gauge installed on the steel chisel to calculate the impact load.

1.0 and αb = 0.55 in Equation (21). The curve with αb = 1.0 over­
Table 3
estimates the attenuation ratio apparently, whereas the curve with αb =
Main parameters of the tested impact source, construction site, and building.
0.55 agrees with the simulated cases very well.
The third configuration simulated the frames with columns, beams Parameters Values
and slabs as well as the material damping ratio. A series of cases with Impact source (concrete breaking by a hydraulic breaker)
varying parameters, including slab thickness s (0–0.25 m), beam cross- Rated energy of the used hydraulic breaker Erated (J) 1000
section dimension hb (0–0.6 m), and equivalent frequency feq Energy transfer efficiency η 40%
Equivalent frequency at impact source f0 (Hz) 1500
(50–1200 Hz) as well as interior, boundary, and corner columns, were Material parameters for ground (thick concrete layer) and building (concrete frame)
simulated to quantify the floor-to-floor attenuation ratio VRf2f. Fig. 14 Young’s modulus: E (GPa) 30
shows that the theoretically predicted VRf2f values generally agree with Density ρ (kg/m3) 2400
the FE simulation results satisfactorily, validating the accuracy of Poisson’s ratio υ 0.2
Material damping ratio ξ for the ground 5%
Equation (13) in all the cases. In particular, the interior, boundary, and
Building structural components
corner columns have beam–slab impedance equal to z = 4zs + 4zb, z = Column cross-sectional dimensions b (m) × c (m) 0.5 × 0.5
2zs + 3zb, and z = zs + 2zb, respectively. Consequently, the interior Slab thickness s (m) 0.15
column exhibits the largest attenuation, whereas the corner column Beam cross-sectional dimensions bb (m) × hb (m) 0.3 × 0.4
exhibits the smallest attenuation. Average floor height h0 (m) 4
Material damping ratio ξc 3%

5. Verification via field measurements


vibration measurement points deployed on the ground surface, column
5.1. Measurement setup base, and the height of a corner column of the tested building. Each
measurement point used a portable measurement system consisting of a
The field measurements of rock excavation-induced vibrations were data acquisition unit (model: NI 9230), an accelerometer (model:
conducted on a selected construction site to verify the proposed for­ 356B18), and a laptop. The vertical vibration velocities were measured
mulas and simulated results. A concrete layer with a thickness of around at a sampling frequency of 5000 Hz. Fig. 15(d) shows a strain gauge
4 m was broken near an existing eight-story concrete frame building. installed on the steel chisel of the hydraulic breaker to calculate the
Fig. 15 shows the hydraulic breaker used for excavation and the impact load shown in Fig. 5. Table 3 lists the major parameters of the

11
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

tested impact source, ground site, and the building, which were also
used in the FE simulations and the proposed prediction formulas.

5.2. Attenuation quantification

Fig. 16 shows the amplitude spectra measured on the free ground


surface at different separation distances (namely, d = 4 m, 8 m, 17 m,
and 27 m) from the impact location. The dominant frequencies are much
higher than the typical frequencies of structures. Due to the faster
attenuation of high-frequency responses, the dominant frequency con­
tents keep decreasing with the increasing measurement distance.
Equation (4) predicted that the dominant frequency decreased from
1030 Hz at d = 4 m to 260 Hz at d = 27 m, which agrees with the
measurement results fairly.
Fig. 17 compares the measured and simulated amplitude spectra at
two column bases with different separation distances d = 2 m and d =
10 m from the excavation point. The frequency contents in the simulated
vertical velocity at the column bases show fair agreement with the
measured results. Both the FE simulation and field measurement show a
Fig. 16. Transient amplitude spectra measured at different distances from the common trend: vibration amplitude and frequency content decrease
impact location. clearly within the propagation from d = 2 m to d = 10 m at column
bases.

(a) (b)
Fig. 17. Comparisons of the measured amplitude spectra with FE simulation results at two column bases with distances (a) d = 2 m and (b) d = 10 m from the
impact location.

(a) (b)
Fig. 18. Variations of (a) measured PPV (mm/s) values, and (b) computed VRb (VdB) with distance d (m) from impact location.

12
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

second. Fig. 20 presents a box-and-whisker plot based on the statistical


results of the attenuation ratios obtained in the field measurement. The
theoretical prediction curve of i × VRf2f according to Equation (13) is
also presented as a straight line in Fig. 20, where the theoretical value of
VRf2f for the tested corner column is predicted as − 2.3 dB for each floor.
The predicted vibration attenuation levels along the column show close
agreement with the field measurement.
In general, the proposed formulas can satisfactorily model the vi­
bration coupling attenuation at column bases and floor-to-floor atten­
uation induced by impact sources such as rock excavation using
hydraulic breakers.
In Fig. 5(a), the hydraulic breaker generated 10 impacts every sec­
ond. By computing the ratio of the peak velocity to RMS velocity in every
second, the crest factor was estimated as CF = 4.5. Accordingly, a simple
conversion of Vrms (VdB) = PPV (VdB) – 13 dB can be performed. The
predicted vibration levels can be further compared with tolerable vi­
bration limits to assess the potential influence on structures, occupants,
Fig. 19. Transient amplitude spectra measured at different floors (column base
and ultraprecision equipment inside buildings. The potentially affected
at d = 10 m). floors of the concerned building with different separation distances from
the impact source can be determined. An example of the vibration
impact assessment is given in the Appendix.
Fig. 18(a) plots the PPV values measured on the free ground surface
and column bases with different separation distances from the impact
6. Conclusions
location. The distances within 24 m are of major interest, where
remarkable vibration impact and apparent vibration attenuation are
This paper proposed a new quantitative assessment framework to
observed. Fig. 18(a) shows two power-law attenuation formulas that
were obtained through regression analyses to represent PPV variations predict building vibrations induced by short-duration impact generated
by hydraulic breakers during rock excavation work. In particular,
with the distance. Subsequently, the attenuation level VRb due to the
coupling effect at column bases was calculated based on two power-law impedance-based prediction formulas were proposed to estimate the
vibration coupling attenuation at column bases and the floor-to-floor
functions. Fig. 18(b) compares the calculated VRb from the field mea­
surement with the theoretical prediction by Equation (9) developed in vibration attenuation inside a building. Their efficacy was verified by
performing systematic FE simulations and field measurement tests. The
this paper. Although both curves show similar variation trends with the
distance, apparent deviation can be observed at close distances while the main conclusions of the current study are summarized as follows:
gap becomes smaller at far distances from the impact location, which
means that the proposed formula led to conservative prediction in this 1) Compared with vibrations induced by traffic, piling, and blasting,
rock excavation using hydraulic breakers generate excessive ground-
case. A likely reason is that wave propagation is more complex in a
stratified ground condition than in the simplified model or FE and structural-borne vibrations with an extremely short duration of
each impact.
simulation.
Fig. 19 shows the amplitude spectra measured at the column points 2) The proposed impedance-based prediction formulas can predict
building vibration levels on any floor of a building at any distance
on different floors (namely, G/F, 2/F, 4/F, 6/F, and 8/F). With the in­
crease of the floor height, the amplitude generally shows an attenuation from the impact source, enabling the direct assessment of vibration
impact on buildings. The comparison with the FE simulation results
trend at high frequencies, although the measured spectra exhibit large
fluctuations with frequency because of more complex structure condi­ and field measurement data verify their satisfactory accuracy.
3) The ground condition, the excitation frequency, and the configura­
tions in field measurement than in FE simulations.
The floor-to-floor attenuation was evaluated by calculating the ve­ tion of structural elements in a multistory frame building influence
the vibration transmission from the ground to the building founda­
locity ratios between the upper floors and the column base in every
tion and from floors to floors.
4) Compared with traditional FE simulation methods, the proposed
impedance-based method presents a simple, computationally effi­
cient framework for vibration predictions, which can be conve­
niently adopted by construction practitioners to assess vibration
impact and design vibration mitigation measures in the construction
planning and design stage.

The current paper presents building vibration predictions with some


ideal assumptions. The complexities arising from stratified ground
conditions, multiple impact sources, and structural irregularities should
be considered in the future to improve the accuracy of the current
assessment framework.

Author statement

Shiguang Wang: Conceptualization, Methodology, Validation,


Formal analysis, Investigation, Data Curation, Writing - Original Draft,
Visualization. Songye Zhu: Conceptualization, Investigation, Writing -
Review & Editing, Supervision, Project administration, Funding
Fig. 20. Calculated velocity attenuation ratios i × VRf2f (VdB). acquisition.

13
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Declaration of competing interest Acknowledgments

The authors declare that they have no known competing financial The authors are grateful for the financial support from the National
interests or personal relationships that could have appeared to influence Key Research and Development Program of China (No.
the work reported in this paper. 2019YFB1600700), and the Hong Kong Polytechnic University (Nos.
ZE2L, BBWJ, ZVX6, and ZJMV). The findings and opinions expressed in
Data availability this paper are solely those of the authors and not necessarily the views of
the sponsors.
Data will be made available on request.

Appendix. An assessment framework for vibration impact induced by rock excavation

This appendix summarizes the computational framework proposed in this paper for predicting building vibrations induced by rock excavation and
assessing the corresponding vibration impact.
The vertical vibrations of the ith floor of a building induced by rock excavation using a hydraulic breaker can be computed as
Lv = Lref + VRb + i × VRf2f , (A1)

where Lv (in VdB) is the peak vibration velocity of the target floor, Lref (in VdB) is the peak ground vibration velocity at a reference point on the free
ground surface, i refers to the floor number and starts from i = 0 for the ground floor, and VRb (in VdB) and VRf2f (in VdB) refer to the coupling
attenuation at column basses and floor-to-floor attenuation, respectively.
The peak ground velocity Lref at a reference point is given by an empirical formula by Wang et al. [21].

[ √̅̅̅̅̅̅̅̅̅̅̅̅ ]
nr
Lref (VdB) = 20 × log 10 kr × ηErated × (d + ar )− (in ​ VdB ​ re: ​ 1 ​ mm/s)
[ √̅̅̅̅̅̅̅̅̅̅̅̅ nr
] (A2)
Lref (VdB) = 20 × log 10 kr × ηErated × (d + ar )− + 92 (in ​ VdB ​ re: ​ 1 ​ μin/s),

where Erated (unit: J) is the rated energy of a given hydraulic breaker, η = 33%–50% is the energy transfer efficiency that is defined as the ratio of the
actual input energy to the rated energy, d (unit: m) is the distance between the vibration source and the point of interest, and three other parameters:
intensity coefficient kr (unit: mm/s/m− n/J0.5), attenuation factor nr, and distance constant ar (unit: m) depend on ground material damping ratio ξ
and/or rock types differentiated by ground shear wave velocity cs (unit: m/s), and they can be determined according to Table A (extracted from Wang
et al. [21]).

Table A
Parameters in Equation (A2) for three typical types of rock mass

Rock type Parameter

kr nr ar

Good quality (cs = 2594 m/s) 33 − 375.6ξ2 + 41.6ξ + 1.1 2.5


Average quality (cs = 1200 m/s) 62 − 375.6ξ2 + 41.6ξ + 1.35 2.0
Poor quality (cs = 473 m/s) 90 − 375.6ξ2 + 41.6ξ + 1.6 1.5

The coupling attenuation VRb is calculated as


⎛ ⎞
⎜ ⎟
⎜ 1 ⎟
VRb (VdB) = 20 × log 10 ⎜ ̅⎟
⎜√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ⎟, (A3)
⎝ 1 + zc + 2 zzc cos(180 − φ1 ) ⎠
| g|
z | g|

where the amplitude and phase angle of the ground impedance zg are
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( 0.5 )− 2
⃒ ⃒
⃒zg ⃒ = Ac ρcs 0.656 + 4.244 ωAc (unit: ​ N/m/s), (A4a)
cs

0.393ωAc 0.5
φ1 = arctan + 90 (unit: ​ deg), (A4b)
cs

where ρ (unit: kg/m3), cs (unit: m/s), Ac (unit: m2), and ω (unit: rad/s) represent the ground density, ground shear wave velocity, foundation area, and
excitation frequency at the foundation, respectively.
The column impedance zc is given by
√̅̅̅̅̅̅̅̅̅
z c = Ac Ec ρ c , (A5)

14
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

where Ec (unit: Pa), ρc (unit: kg/m3), and Ac (unit: m2) are Young’s modulus, mass density, and cross-sectional area of the column, respectively.
The floor-to-floor attenuation VRf2f is the sum of two effects.
1 ωξ h0
VRf2f = VRf2f− 1 + VRf2f− 2 ̅ − 8.68 × c ,
= 20 × log 10 √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 (A6)
|z| |z|
cp
1 + 2zc + 2 2zc cos φ2

where the first term VRf2f-1 (VdB) refers to the floor-to-floor velocity amplitude ratio due to energy division at the column–beam–slab connection for
each floor, and z is the sum of the impedance contributions from the floor slabs and beams. Assume zs is the impedance of a floor slab in one quadrant,
and zb is the impedance of a beam on one side. The value of z depends on the numbers of the beams and slabs connected to the column: z = 4zs + 4zb for
an interior column, z = 2zs + 3zb for a boundary column, and z = zs + 2zb for a corner column. φ2 is defined as the phase angle of the total impedance z.
The diving point slab impedance zs in one quadrant is expressed as
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅

zs = αs ⋅2s2 ( s s ), (A7)
12 1 − υ2s

where Es (unit: Pa), ρs (unit: kg/m3), υs, and s (unit: m) are Young’s modulus, mass density, Poisson’s ratio, and thickness of the slab, respectively. The
modification coefficient αs = 1.3 is suggested.
The diving point beam impedance zb on one side is expressed as
αb
(A8)
3 1 3 1
zb = bb hb 2 Eb 4 ρb 4 ω2 (1 + j),
121/4

where Eb (unit: Pa), ρb (unit: kg/m3), ω (unit: rad/s), and bb (unit: m) × hb (unit: m) are the Young’s modulus, mass density, excitation frequency, and
cross-section dimensions (horizontal × vertical) of the beam, respectively. The modification coefficient αb = 0.55 is suggested.
The second term VRf2f-2 (VdB) in Equation (A6) refers to the attenuation between two adjacent floors (i.e., from the ith floor to (i+1)th floor)
caused by material damping. cp (unit: m/s), ω (unit: rad/s), ξc, and h0 (unit: m) are the compressive wave velocity, dominant excitation frequency,
material damping ratio in the column, and the floor height, respectively.
The equivalent frequency ω of the impact-induced vibration used in Equation (A4a), Equation (A6), and Equation (A8) can be estimated by
( )
2ω π cs
ω = 0 arctan , (A9)
π 2ω0 ξd

where ω0 (unit: rad/s), cs (unit: m/s), and ξ are the dominant frequency at impact source, ground shear wave velocity, and ground damping ratio,
respectively; d (unit: m) is the separation distance from the impact source. The value of ω0 can be determined from field measurement results, and ω0
= 9425 rad/s (i.e., f0 = 1500 Hz) is suggested for the impact load induced by a hydraulic breaker.
The RMS velocity level Lrms (in VdB) at the target floor can be estimated as
Lrms (VdB) ​ = ​ Lv − 20log 10 CF, (A10)

where CF is a crest factor defined as the ratio between the peak velocity and the RMS velocity. CF = 4.5 is suggested for impact vibrations induced by a
hydraulic breaker generating 10 impacts in every second.
Subsequently, the predicted vibration levels Lrms (VdB) can be compared with the tolerable vibration limits considering occupants and ultra­
precision equipment to evaluate the potential vibration impact on them.
The vibration impact generated by concrete plate excavation on the eight-story concrete building was assessed as an example based on the in­
formation provided in Table 3. Considering two separation distances d = 2 m and d = 24 m from the impact souce to the column bases, the reference
ground velocity Lref decreased from 119.1 dB (re 1 μ-in/s) at 2 m to 84.6 dB at 24 m, and VRb increased from − 5.0 dB at 2 m to − 0.7 dB at 24 m, The
variation of VRf2f depends on the column types: VRf2f increased from − 3.7 dB to − 1.4 dB for corner columns, from − 4.5 dB to − 1.9 dB for boundary
columns, and from − 5.4 dB to − 2.6 dB for interior columns, when separation distance d increased from 2 m to 24 m.
The peak vibration velocity level Lv (VdB) on any floor of a building was then calculated by Equation (A1), and RMS velocity level Lrms (VdB) was
estimated through the simple conversion of Equation (A10). Fig. A1 shows the predicted Lrms (VdB re 1 μ-in/s) on different floors for three types of
columns (interior columns, boundary columns, and corner columns) at a distance d = 10 m from the impact source. Allowable vibration impact levels
Lrms (VdB re: 1 μ-in/s) suggested by FTA [8] for three categories of building use, namely, 65 dB for the operation of sensitive equipment, 72 dB for
normally sleeping people, and 75 dB for primarily daytime use, were adopted to evaluate the predicted vibration velocity level. The hollow circle
marked in Fig. A1 indicates the floor below which the corresponding category of building use is expected to be affected. Accordingly, the highest floor
number to be affected at various distances from the impact source can be determined.

15
S. Wang and S. Zhu Soil Dynamics and Earthquake Engineering 163 (2022) 107508

Fig A1. Variations of predicted Lrms (VdB) with the floor number i at a distance d =10 m from the impact source.

References [15] Chua KH, Balendra T, Lo KW. Groundborne vibrations due to trains in tunnels.
Earthq Eng Struct Dynam 1992;21(5):445–60. https://doi.org/10.1002/
eqe.4290210505.
[1] Caltrans. The transportation and construction induced vibration guidance manual.
[16] Lopes P, Alves Costa P, Ferraz M, Calçada R, Silva Cardoso A. Numerical modeling
Sacramento: Caltrans; 2013.
of vibrations induced by railway traffic in tunnels: from the source to the nearby
[2] Sanayei M, Zhao N, Maurya P, Moore JA, Zapfe JA, Hines EM. Prediction and
buildings. Soil Dynam Earthq Eng 2014;61:269–85. https://doi.org/10.1016/j.
mitigation of building floor vibrations using a blocking floor. J Struct Eng 2012;
soildyn.2014.02.013.
138(10):1181–92. https://doi.org/10.1061/(ASCE)ST.1943-541X.0000557.
[17] Yang J, Zhu S, Zhai W, et al. Prediction and mitigation of train-induced vibrations
[3] Athanasopoulos GA, Pelekis PC. Ground vibrations from sheetpile driving in urban
of large-scale building constructed on subway tunnel. Sci Total Environ 2019;668:
environment: measurements, analysis and effects on buildings and occupants. Soil
485–99. https://doi.org/10.1016/j.scitotenv.2019.02.397.
Dynam Earthq Eng 2000;19(5):371–87. https://doi.org/10.1016/S0267-7261(00)
[18] Auersch L. Dynamic stiffness of foundations on inhomogeneous soils for a realistic
00008-7.
prediction of vertical building resonance. J Geotech Geoenviron Eng 2008;134(3):
[4] Dowding CH. Construction vibrations. Englewood Cliffs: Prentice Hall; 1996.
328–40. https://doi.org/10.1061/(ASCE)1090-0241(2008)134:3(328).
[5] Oriard LL, Richardson TL, Akins KP. Observed high-rise building response to
[19] Sanayei M, Moore JA, Brett CR. Measurement and prediction of train-induced
construction blast vibrations. In: Vibration problems in geotechnical engineering;
vibrations in a full-scale building. Eng Struct 2014;77:119–28. https://doi.org/
1985 [Detroit, Michigan, United States].
10.1016/j.engstruct.2014.07.033.
[6] Xia H, Chen J, Wei P, Xia C, De Roeck G, Degrande G. Experimental investigation of
[20] Zou C, Moore JA, Sanayei M, Wang Y. Impedance model for estimating train-
railway train-induced vibrations of surrounding ground and a nearby multi-story
induced building vibrations. Eng Struct 2018;172:739–50. https://doi.org/
building. Earthq Eng Eng Vib 2009;8(1):137–48. https://doi.org/10.1007/s11803-
10.1016/j.engstruct.2018.06.032.
009-8101-0.
[21] Wang S, Zhu S, Yuen PL. Assessment of ground-borne vibration impact on nearby
[7] Zou C, Wang Y, Wang P, Guo J. Measurement of ground and nearby building
underground facilities induced by ground surface excavation. J Construct Eng
vibration and noise induced by trains in a metro depot. Sci Total Environ 2015;536:
Manag 2021;147(7):04021071. https://doi.org/10.1061/(ASCE)CO.1943-
761–73. https://doi.org/10.1016/j.scitotenv.2015.07.123.
7862.0002065.
[8] FTA (Federal Transit Administration). Transit noise and vibration impact
[22] Massarsch KR. Static and dynamic soil displacements caused by pile driving. In:
assessment manual. Washington: FTA; 2018.
Proc. Of keynote lecture, 4th int. Conf. On the application of stress wave theory to
[9] Sanayei M, Maurya P, Moore JA. Measurement of building foundation and ground-
piles; 1992 [Rotterdam, Netherlands].
borne vibrations due to surface trains and subways. Eng Struct 2013;53:102–11.
[23] Telford WM, Geldart LP, Sheriff RE, Keys DA. Applied geophysics. New York:
https://doi.org/10.1016/j.engstruct.2013.03.038.
Cambridge Univ. Press; 1990.
[10] Auersch L. Response to harmonic wave excitation of finite or infinite elastic plates
[24] Thomson WT, Kobori T. Dynamical compliance of rectangular foundations on an
on a homogeneous or layered half-space. Comput Geotech 2013;51:50–9. https://
elastic half-space. J Appl Mech 1963;30(4):579–84. https://doi.org/10.1115/
doi.org/10.1016/j.compgeo.2013.02.001.
1.3636622.
[11] Sanitate G, Talbot JP. On the plate-like and layer-like response of slab foundations
[25] Richart FE, Hall JR, Woods RD. Vibrations of soils and foundations. Englewood
to ground-borne vibration. Comput Geotech 2019;114:103141. https://doi.org/
Cliffs, NJ: Prentice Hall; 1970.
10.1016/j.compgeo.2019.103141.
[26] Cremer L, Heckl M, Ungar EE. Structure borne sound. Berlin: Springer; 1988.
[12] Dolšek M, Fajfar P. Simplified non-linear seismic analysis of infilled reinforced
[27] Kirchhoff GR. Uber das gleichgewicht und bewegung einer elastischen scheibe.
concrete frames. Earthq Eng Struct Dynam 2005;34(1):49–66. https://doi.org/
J für die Reine Angewandte Math (Crelle’s J) 2009;1850(40):51–88. https://doi.
10.1002/eqe.411.
org/10.1515/crll.1850.40.51.
[13] Hong XJ, Zhu S, Xu YL. Three-dimensional vibration control of high-tech facilities
[28] Fahy FJ, Gardonio P. Sound and structural vibration: radiation, transmission and
against earthquakes and microvibration using hybrid platform. Earthq Eng Struct
response. Amsterdam: Elsevier; 2007.
Dynam 2010;39(6):615–34. https://doi.org/10.1002/eqe.960.
[29] Hoek E, Brown ET. Practical estimates of rock mass strength. Int J Rock Mech Min
[14] Zhu S, Shi X, Leung R, et al. Impact of construction-induced vibration on vibration
Sci 1997;34(8):1165–86. https://doi.org/10.1016/S1365-1609(97)80069-X.
sensitive medical equipment: a case study. Adv Struct Eng 2014;17(6):907–20.
[30] Auersch L. Simple and fast prediction of train-induced track forces, ground and
10.1260%2F1369-4332.17.6.907.
building vibrations. Railw Eng Sci 2020;28(3):232–50. https://doi.org/10.1007/
s40534-020-00218-7.

16

You might also like