You are on page 1of 34

On Comparing the Characteristics of Rotational and Appropriate Translational Ground

Motions

Gopala Krishna Rodda1 and Dhiman Basu2

Abstract

Rotational ground motion may contribute significantly in the response of certain structures. To date,
these components are not measured by the accelerographs deployed in free-field owing to the non-
availability of appropriate instruments and hence, their effects are generally ignored in seismic design.
Indirect methods that enable extraction of these components from the three-component translational
recording have been reported in literature to understand more about its consequence on the response of
structures. This paper explores the possible correlation of extracted rotational and recorded translational
ground motions. An appropriate translational component is defined such that time derivative of which
is closely related to the associated rotational component. Several quantitative descriptions are included
in this investigation including the response spectra, auto-spectral density and relative energy build up.
A new parametric representation of auto-spectral density is presented that enables extraction of several
meaningful insights. A new definition of apparent wave velocity is proposed.

Keywords: Rotational ground motion; appropriate translational component; parametric auto-spectral


density; relative energy build up; pulse-like ground motion.

Introduction

At the advent of strong motion network, translational ground motion database is now available almost
all over the world in order to facilitate seismic design. It is impossible to predict a future earthquake in
terms of the expected time series, but essential to know its probable characteristics for the seismic
design. Phase uncertainty is an example that precludes a ground motion to be predicted in terms of time
series. Even though complete time history information is required for the deterministic analysis and
design of a structure, some specific characteristics of the ground motion only affect the structural
responses. Several ground motion parameters are reported in literature for the quantitative descriptions
including the Peak Ground Acceleration (PGA), Peak Ground Velocity (PGV), Peak Ground
Displacement (PGD), frequency content of the motion and strong duration of the event, cumulative
relative input energy build up or its spectral measures etc. Apart from three translational components,
three rotational components are also required for the complete description of a seismic event. But, the
rotational components of a ground motion are currently not included in the response history analysis of
structures because they are not measured directly by the accelerographs deployed in the free-field.
Accordingly, their effects on the response of structures are largely unknown. A number of articles have
been reported on the effect and importance of rotational components of ground motion on structural
response.

1
PhD candidate, Department of Civil Engineering, Indian Institute of Technology Gandhinagar,
Ahmedabad, India-382424
2
Corresponding Author, Assistant Professor, Department of Civil Engineering, Indian Institute of
Technology Gandhinagar, Ahmedabad, India-382424, Email: dbasu@iitgn.ac.in

Wolf et al. (1983) discussed the effect of rocking excitation on a base-isolated nuclear power plant.
Politopoulos (2010) identified the excitation of the rocking mode in a base-isolated building due to
rocking excitation. Both studies were based on simplified assumptions including horizontally
propagating waves in bedrock and vertically propagating shear waves. Zembaty and Boffi (1994) and
Zembaty (2009a) identified the contribution of rocking motion to bending moments along the height of
a 160 m tall tower using horizontal and rocking spectra computed based on Eurocode 8 (EC8.1, 2005;
EC8.6, 2005). De La Llera and Chopra (1994) studied the effect of torsional ground motion, which was
reported more critical in a near symmetric building. Recently, Basu et al. (2014) studied the accidental
eccentricity contributed from the torsional ground motion and reported that shifting of the centre of
mass, as commonly used in accounting for accidental torsion, is not suitable, especially, for response
history analysis. Also proposed was a possible alternative with appropriate illustration. Basu and Giri
(2015) further extended the approach to multistory buildings subjected to bidirectional seismic
excitation. Falamarz-Sheikhabadi et al (2013) reported that principal plane of the rotational motion
could be different from that of the translational motion. Estimated rotational components led to spectral
density function of the rocking components three times larger than those considered for the rotational
loading of structures in the worst-case scenario. Sheikhabadi (2014) proposed simplified relation to
account for the rotational ground motion in seismic design. Despite these efforts, consensus on the
importance of rotational components of ground motion on structural response does not exist. Basu et
al. (2015) investigated the effect of rotational (torsional and / or rocking) ground motion on the response
of sample structures that includes a tall chimney and base-isolated building.

A lack of consensus can be attributed, in part, to the use of different interpretations of rotation, namely,
the free-field or point rotation, chord rotation and averaged rotation. Free-field rotation is defined as the
spatial derivative of the instantaneous displacement field; chord rotation is the ratio of the difference in
displacements between two closely spaced, adjacent stations measured along a direction normal to the
connecting line to the separation distance along the connecting line; and average rotation is the average
(or weighted average) of chord rotations of several station pairs with one common station. One may
argue about the choice of one of these three rotations for response-history analysis. Point rotation
preserves the frequency content of the ground motion whereas the use of chord rotation or average
rotation will filter or eliminate high frequency components. A rigid foundation will filter or eliminate
high frequency components of free-field motion although foundations are never infinitely stiff.
Accordingly, average rotation may be the best choice because chord rotation is usually sensitive to
separation distance (Laouami and Labbe, 2002) and point rotation ignores the presence of a foundation.
Regardless of the choice of rotation, the first step is to generate free-field rotational ground motions,
which can then be modified, on a case-by-case basis, to account for foundation dimensions and stiffness.
Therefore, spatial variability of translational and, especially, rotational ground motion is required to
completely specify the appropriate ground motion inputs. Seismic array/ dense array, a cluster of closely
spaced recording stations, provides a unique opportunity to study the spatial variability. A number of
researchers have studied the characteristics of translational ground motion, its spatial variability and
possible extraction of associated rotational components. Among these, the studies relevant to the
context of present paper have been summarized below.

On Characteristics of Translational Ground Motion

Some of the ground motion characteristics, for example, PGA, PGV/PGA and strong motion duration
are closely correlated with the energy dissipated by a SDOF system (Haluk et al.,1995). Energy
dissipated during an earthquake represents the damage potential of a ground motion. These parameters
were studied to characterise the damage potential and PGA does not show well correlation with the
maximum internal force, especially, in the events with short duration. An energy dissipation index was
proposed as a measure of the damage potential: [An elastic SDOF subjected to the chosen ground
motion was analysed for the energy dissipation due to viscous damping at the end of the event. An
equivalent velocity was next computed based on the analogy with kinetic energy and plotted in the form
of a spectrum. The energy dissipation index was then calculated by averaging the equivalent spectral
velocity over a certain range of time period (usually 0 to 4 seconds).

Ellen et al. (1998) characterized the frequency content of an earthquake ground motion with a single
parameter. Three parameters were examined as a possible candidate, namely, the mean period ( Tm ),
predominant period ( T pr ) and smoothed spectral predominant period ( To ). An empirical model
describing the correlation of these frequency parameters with magnitude ( M ) and epicentral distance
( R ) was developed using a large database of ground motion. Linear regression analysis was performed
to get the regression coefficients. The results indicate that the parameter T pr (period at which

2
acceleration spectra reaches the maxima) has the largest uncertainty in its prediction. The parameter Tm
(weighted average of time period with square of Fourier amplitudes being the weights) is found to be
the best simplified frequency content parameter and it can be reliably estimated.

Tso et al (1992) investigated the significance of the parameter PGA/PGV ratio as an indicator of the
dynamic characteristics of earthquake ground motions. A set of 45 ground motions were selected for
analysis, which were further categorized into three cases, namely, low, intermediate and high depending
upon the ratio PGA/PGV. Average value was calculated in each category and following observations
were made: i) strong negative correlation between the strong motion duration and PGA/PGV ratio; and
ii) similar trend for the Housner’s spectral intensity. Peak ground acceleration (high frequency)
attenuates rapidly with the distance when compared to the peak ground velocity (moderate frequency).
Therefore, ground motions near an earthquake source have higher PGA/PGV ratios than that recorded
at a large distance from the source for the same earthquake.

Based on the analysis of strong motion events recorded at the SMART1 array, Harichandran (1991)
reported that the spatial variability of free-field translational ground motion could be significant within
the dimension of a typical large engineered structure. Lagged coherency is generally used as a measure
of the spatial variability. Note that the coherency can be obtained first by normalizing the smoothed
cross-spectral density and lagged coherency is given by its absolute value. It was reported to be reducing
with the increasing separation distance and frequency. Interestingly, the attenuation of lagged coherency
is not overly direction sensitive, which has led to a considerable amount of research in the last two
decades on modelling the spatially varying earthquake ground motion (translational) and, on
determining its effect on the seismic response of rigid mat foundations and large structures such as
bridges, pipelines, dams etc.

Several other studies have been reported discussing relative merits and demerits of possible quantitative
descriptions of translational ground motion and a comprehensive review of which is beyond the scope
of the present context.

On Extraction of Rotational Ground Motion

Rotational components ground motion have been the interest of many researchers and the first study
probably dates back to late ’60s. Newmark (1969) showed that a symmetrical building could be excited
by the torsional component of ground motion on the basis of a travelling wave propagating with a
constant velocity. Hart et al. (1975) attributed the torsional response of most high-rise buildings during
the 1971 San Fernando earthquake to the rotational components of the ground motion.

Defining the vertical plane comprising of the recording station and the epicentre as the principal plane,
and its projection on the horizontal plane as the principal axis, Penzien and Watabe (1975) showed most
of the energy travels to the station on this principal plane and the three components along and normal
to the principal plane are uncorrelated. In the absence of surface waves, this idealization simplifies the
problem. The contribution of the SH wave can then be readily identified as the resultant of the
components of the recorded horizontal ground motions normal to the principal plane. P and SV waves
contribute to the ground motion along the principal axis and in the vertical direction. Bolt (1988)
provides information on these body waves. Trifunac (1982) derived the Fourier amplitude spectra of
rotational components of ground motion from the translational components assuming the incidence of
a single type of body wave (P or SH or SV). A procedure for generating the strong motion translational
accelerograms from a given Fourier amplitude spectra and strong motion duration (Wong and Trifunac,
1979) was extended to torsional accelerograms (Lee and Trifunac, 1985). Bouchon and Aki (1982)
analytically studied the characteristics of the propagation of earthquake-induced dynamic strain and
rotational fields for two source mechanisms: strike-slip and dip-slip faults. The displacement field
associated with each source mechanism was computed using a discrete wave-number method. Bouchon
and Aki (1977) proposed this method first in two spatial dimensions. Bouchon (1979) extended the
method to three spatial dimensions and then applied the method to strike-slip and reverse fault
earthquake sources (Bouchon 1980a, b). They first decomposed the contributions from the surface and

3
body waves using a technique described in Sugito et al. (1984). The time- and frequency-separation
parameters required for this decomposition are computed from the evolutionary spectra discussed in
Kameda (1975). The body wave contributions were then decomposed into P and S waves using a theory
developed by Haskel (1953). Hao (1998) investigated the characteristics of torsional ground motion and
its relation to the corresponding translational motions in terms of their respective power spectral
densities (PSD) and response spectra. Li et al. (2002) estimated the frequency-dependent velocities of
the surface waves, which can improve the estimation of rotational components from far-field recordings
where the contributions of the body waves are close to zero. Falamarz-Sheikhabadi et al (2012)
estimated rotational components based on small displacement gradient, only shear wave incidence and
apparent wave velocity (frequency independent). A number of researchers have used the definitions
offered by Penzien and Watabe (1975) to extract rotational time series from measured translational
recordings at a single station, including Trifunac (1982), Lee and Trifunac (1985), Lee and Trifunac
(1987), Castellani and Boffi (1986), Castellani and Boffi (1989), Gomberg (1997), Zembaty (2009), Li
et al. (2004) and Basu et al. (2012). The available single station procedures involve a number of
assumptions, including plane wave propagation, existence of a principal plane, lateral homogeneity of
the soil medium, a frequency-dependent angle of incidence, the effect of dispersion, and the
indeterminacy involved in the deconstruction of the recorded translational time series to contributions
from different types of body and surface waves. All the studies reported above relate the rotational
ground motion (or it’s spectral representation) with the three-component translational ground motions
(or respective spectral representations) recorded at a single point and hence, collectively referred to the
Single Station Procedure (SSP).

An alternative to the single station procedure involves the use of data from a number of closely spaced,
spatially distributed stations: the so-called dense array. These multiple station procedures (MSP), have
formed the basis of studies to estimate rotational components of ground motion from recorded
translational acceleration time series. Niazi (1986) estimated rotational motions from the data recorded
in a linear array, the El Centro Differential Array (ECDA) in Southern California. The procedure
involves fitting a best-fit straight line across the array at every instant of time. The Geodetic Method
(GM) of Spudich et al. (1995) can be considered as a significant expansion of the work of Niazi. Stations
were distributed in a three-dimensional space and the relative displacement between any station pair
(with a common station, called the reference station) was expressed in terms of a displacement -gradient
matrix, which was assumed to be constant and small over the footprint of the array. Basu et al. (2013)
noted that: (1) the three constraint equations used to relate the displacement gradient parameters in GM
are valid only at the free surface, and thus the use of recordings at depth should be avoided; (2) when
using recordings at surface stations only, the procedure may be thought of as calculating the rotational
components induced in a two dimensional (instead of line-element) rigid-foundation whose geometry
is defined by the spatial distribution of the stations; and (3) the rotational component computed using
the GM does not necessarily characterize the free-field component. The rotational component computed
using the GM does not necessarily characterize the free-field component.

Using data recorded at surface stations of the LSST array, Lotung (Taiwan), Laouami and Labbe (2002)
computed the torsional component for each station pair by first rotating the recorded horizontal motions
along the direction normal to the line connecting the stations and then dividing the differential motion
by the separation distance. Assuming stationarity and identifying the strong-motion phase for each
station pair, the spatial variability was then studied by defining a normalizing parameter that is
proportional to the separation distance and the ratio of the standard deviation of the torsional motion to
the average standard deviation of the horizontal motions used in calculating the torsional motion.
Researchers have used the geodetic method of Spudich et al. (1995) to estimate rotational components
of ground motion (e.g., Bodin et al., 1997; Ghayamghamian and Nouri, 2007). Estimates of rotational
components and dynamic strains computed using the geodetic method and data reconstructed from a
single recording station have been compared (Gomberg, 1997; Castellani and Boffi, 1989; Singh et al.,
1997) and the reported differences were relatively small. Bodin et al. (1997) concluded that an array
dimension of one-quarter of the wavelength of the dominant mode would provide an estimate of the
gradient that was within 10% of the true value by comparing the numerical gradient with the actual
gradient per Lomnitz (1997). Basu et al. (2013) reported that this comparison of the gradient is true at

4
a single station and based on a single station pair. In contrast, the gradient returned by the geodetic
method is, in some sense, the weighted average of those computed using several station pairs sharing
one common station. Basu et al. (2013) reported a unified procedure capable of extracting the free-field
rotational time series from the three-component strong motion data recorded in a dense array. The
unified procedure was built on the seminal work of Spudich et al. (1995). Only surface station data were
used and that enabled the entire process treating the recorded translational components in an uncoupled
way.

In general, the SSP is unaware of the spatial variability of the translational components and calculates
the free-field or point rotation under the assumption of plane wave propagation etc. On the other hand,
MSP uses the dense array recording and reports the average chord rotation, which fails to capture the
high frequency components of the recorded ground motion. In order to overcome this issue, Basu et al.
(2015) developed a framework that is same as the MSP in principle but carried out in frequency domain.
Resulting procedure enables capturing of high frequency components while using the dense array data.

General Inferences and Objectives of Present Study

Rotational ground motion may contribute significantly to the response of certain structures and is not
currently recorded from the accelerographs deployed in the free field due to lack of appropriate
instruments. Therefore, indirect method, e.g., SSP and MSP are used to extract the rotational ground
motion from the recorded three component translational acceleration data. Among these, SSP is a more
common choice as the dense array recording is rare when compared to the recording stations in strong
motion network across the globe. In general, plane wave propagation theory assumed in SSP enables to
relate the rotational motion (spatial derivative of the displacement field) with the point velocity measure
(temporal derivative) through phase velocity as a function of frequency, incident angle etc. Since wave
propagation characteristics differs with frequency, resulting rotational time series is not related with the
temporal derivative time series in a simple way.

The objective of this paper is to explore the possible correlation of rotational ground motion
characteristics with that of the translational motion. SSP developed in Basu et al. (2012a) is used as the
basis of extracting the rotational ground motion. Towards this, the paper will seek existence of an
appropriate translational motion such that time derivative of which is closely related to the rotational
ground motion (torsional and rocking). Several characteristics are used in this investigations such as,
5% damped response spectrum, auto spectral density, relative energy build up, energy spectrum, strong
motion duration etc. Apparent wave velocity is more commonly used in extracting the rotational ground
motion and the present paper recommends a new definition for it. Simplified method proposed in this
paper is aimed to develop an alternative SSP and objective is not to pursue MSP which requires the data
recorded in a dense seismic array. Such array is not available in most part of the world.

Description of seismic array and events considered

The Lotung-LSST (LLSST) site is a part of the much larger SMART1 array. Figure 1 shows the surface
stations of the LSST array to be considered in this paper. All fifteen free-surface accelerometers at the
LSST are positioned along three arms at approximately 120 degree intervals. Each arm extends for
about 50 m and the spacing between the surface stations varies from 3m to 90 m. Each arm contains
five stations that are designated here as 1 through 5, starting at the centre of the array. For example,
FA1_1 and FA2_5 denote the innermost and outermost stations located on arm 1 and 2, respectively.
Further details on the site characteristics, instrumentation and recorded seismic event may be obtained
from the URL http://www.earth.sinica.edu.tw/~smdmc/llsst/llsst.htm. The average wave velocities at
the surface layer of the recording site are: 140 m/sec and 595 m/sec for the S and P waves, respectively
(Wen and Yeh, 1984).

Three strong motions events recorded at LSST array are considered for analysis in this paper. Brief
description of each event is presented in Table 1. Detailed description of the recorded data are provided
later in this paper. Note that Event-3 may have some near-field effect as the epicentral distance is

5
approximately 20kM. Only surface stations are considered in the analysis and out of 15, usually, 10-14
actually functioned during the events. Hence, number of surface stations analysed varies from one event
to another.

Recorded and Principal Directions

Ground motion is recorded in LSST array along the East-West (EW), North-South (NS) and vertical
directions. Recorded horizontal accelerations (EW and NS) are often rotated along and normal to the
principal plane to facilitate extraction of rotational components. These two rotated components, i.e.
along and normal to the principal planes, are denoted in this paper as a g 1 and a g 3 , respectively; for
completeness, vertical acceleration is denoted as a g 2 .

Extraction of rotational components

Event-1 is considered first and the time series of rotational components at any surface station are
extracted from the three-component translational data recorded at the same station. Detailed explanation
of method of extraction of rotational components is given in Basu et al. (2012). Rotational components
include the torsional ground motion (rotation about the vertical axis) and rocking acceleration on the
principal plane. Note that rocking component on a plane normal to the principal plane is zero. Similar
analysis is carried out at all the surface stations functioned during the event. Finally, Event-2 and Event-
3 are also analysed for the rotational time series.

Spectral characteristics of translational and rotational components

Five percent damped response spectra are computed for the translational and rotational time series.
Translational time series includes the vertical components and also the rotated a g 1 and a g 3 components.
Similarly, rotational spectra include both torsional and rocking on the principal plane. The parameters
used for the purpose of comparison regardless of ground motion components, are Peak ground
acceleration (PGA, spectral ordinate at zero time period) and the measure of similarity of spectral
shapes. Figure 2a presents the response spectrum of a g 1 at station FA1_1 in Event-1. A sample
rotational spectrum (torsional) is included in Figure 2b (the same station and the same event).

PGA is tabulated in Table 2 for the recorded ground motions in Event-1. The representation provides
an idea of the expected variation of structural response on account of the spatial variability of recorded
ground motion. Table 2, also, presents the similar information for the extracted rotational motions. Also
included are the spectral characteristics of derivative of the a g 1 , a g 2 and a g 3 . Similar data for the other
two events are not presented here for brevity.

Existence of appropriate translational motion

By, definition, rotational motion is a spatial derivative of appropriate translational motion, which can
be converted to time derivative of that translational motion with due scaling through apparent wave
velocity. An appropriate translational component is defined such that time derivative of which is closely
related to the associated rotational component.

Torsional Ground Motion

PGA of torsional component and that of translational components (from Table 2) have been observed
and no correlation is observed. Events -2 and -3 also exhibit the similar trend. Similar comparison is re-
exercised but replacing the spectral characteristics of translational components by that of its time
derivatives. A strong correlation has been observed between PGAs of torsional component and a g 3 .

6
Results are presented in Figure 3 for all the three events. PGAs in a g 3 and torsional motion are seen to
be correlated in most stations.

PGA, which is just a point on response spectra, alone cannot determine the relation between a g 3 and
torsional motion. Hence, distance correlation has been chosen to measure the resemblance between the
two response spectra. Correlation being zero does not imply independence while distance correlation
being zero does imply independence.

Let (Xk, Yk), k= 1, 2, ..., n be a statistical sample from a pair of real valued random variables (X, Y).
First, compute all pairwise distances.

a j ,k  X j  X k
(1)
b j ,k  Y j  Yk , j , k  1, 2, ...., n

where || ⋅ || denotes Euclidean norm. That is, compute the n by n distance matrices (aj,k) and (bj,k). Then
take all doubly centered distances.

Aj ,k  a j ,k  a j .  a.k  a..
(2)
B j ,k  b j ,k  b j .  b.k  b..

where a j . is the jth row mean, a.k is the kth column mean, and a.. is the grand mean of the distance
matrix of the X sample. The notation is similar for the b values. The squared sample distance
covariance is simply the arithmetic average of the products Aj,k Bj,k.
n
1
dCovn2 ( X , Y ) 
n2
A
j , k 1
j ,k B j ,k (3)

The distance correlation of two random variables is obtained by dividing their distance covariance by
the product of their distance standard deviations.
The response spectra obtained is generally sporadic in nature. Hence before comparing the spectral
shape, it is suggested to smoothen the spectra. Hamming window is used to carry out smoothening in
this paper. The shape of the Hamming window is similar to that of a cosine wave. This method considers
an L -point symmetric window at each data point. The window length is L  2M  1 , where M is the
number of hamming points considered around each point and M  m  M is any particular hamming
point. The weight w  m at each hamming point is given as

  (m  M ) 
0.54  0.46cos M 
w(m)  , For m  - M , - M  1, , M –1, M (4)
1.08M
The smoothened value at the data point is the weighted average of all the surrounding points within the
window:

j  M
SAsmooth (Ti )   SA(T  jT )w( j)
j  M
i (5)

7
The response spectra of the ground motion (both a g 3 and torsional motion) is first smoothed with a 20
point hamming window to get smoothed response spectrum. Distance correlation obtained between the
pair of them for all events has been reported in Table 3. Correlation is found to be close to at all stations
for all events. Also, difference between the actual response spectra and the smoothened response spectra
is termed as deviation spectra. Distance correlation is then calculated between the square of the
deviation spectra of rotational motion and appropriate translational motion derivative and the range
correlation coefficients obtained were 0.74 to 0.95. Hence, a relationship of direct proportionality
possibly exists between the torsional acceleration and a g 3 .

Despite the dependency through frequency dependent incident angle, torsional acceleration (time series)
can be well characterized through a g 3 and hence, a g 3 component may be identified as the appropriate
translational motion for torsional component.

Rocking Ground Motion

Even though torsional motion is fully described by the translational motion a g 3 , rocking motion is
contributed from both a g 1 and a g 2 , similar comparison (as above) is carried out with respect to both
the translational components, but no correlation is obtained (Table 2). Similar comparison is also
extended with respect to the derivative of the translational accelerations, i.e. a g 1 and a g 2 (Table 2).
Once again, no correlation between rotational and translational components is evident.

Appropriate translational component for rocking motion needs an approach from first principles. The
simplified procedure is based on the assumption of near vertical incidence of P and S-waves, regardless
of the harmonics. Since the density of soil increases with depth, near vertical incidence is a reasonable
assumption, especially at some distance away from the hypocenter. Contribution of P-wave in the
estimated rocking motion in rth harmonic on the principal plane may be expressed as (Basu et al., 2012a)
p ..
sin  0 p p . 2 2sin 2  0 p p .
rxx t 1 2 r ag 2 t 1 r ag 1 t (6)
cL
2cL  2 sin 2  0 p 2

Here,  0 p denotes the angle of incidence, and  cL cT is the ratio of P to S wave velocity. Further,
first and second parts of the equation estimate the same rocking component from the derivative of the
vertical and horizontal motions, respectively. Assuming the same angle of incidence, regardless of the
harmonics, and using the superposition of all harmonics, one may show
p ..
sin  0 p p .  2 2sin 2  0 p p .
 x1x2 t ag 2 t 1
ag 1 t (7)
cL
2cL  2
sin  0 p
2 2

Note that the subscript r is dropped while the superscript p is retained to emphasis the P-wave
contribution. Further, for near vertical incidence (  op  0 ), one may prove that (Basu et al., 2012)
p p
ag2 t ag1 t 0 (8)

SV-wave field, on the other hand, depends on the angle of incidence with respect to critical. However,
the present interest is on the near vertical incidence and hence, the angle of incidence will be always
less than the critical. For such cases, contribution of SV wave in the rocking motion on the principal
plane in r th harmonic may be expressed as

8
0.5
sin 0 s s . 2sin 2 0 s 1  sin 0 s
s .. 2 2 s .
r xx t r ag 2 t r ag 1 t (9)
 cT 1 2sin 2 0 s
1 2
cT

Assumption of constant (near vertical) angle of incidence and superposition of all harmonics lead to the
rocking motion as
0.5
sin 0 s s . 2sin 2 0 s 1  sin 0 s
s .. 2 2 s .
xx t ag 2 t ag 1 t (10)
 cT 1 2sin 2 0 s
1 2
cT

Note the presence of superscript s to signify the contribution from the SV-wave. Further, for near
vertical incidence ( os  0 ), one may prove that (Basu et al., 2012)
s s
ag1 t ag 2 t 0 (11)

Based on near vertical but simultaneous incidence of P and SV waves, rocking component can
expressed as
0.5
sin 0 p 2sin 2 0 s 1  sin 0 s
p 2 2 s
.. . .
xx t ag 2 t ag 1 t (12)
 cT 1 2sin 2 0 s
1 2
cL

Note that total rocking is expressed as the superposition of P and SV wave contributions. Further, based
on Eq (8) and Eq (11), assuming
p
ag2 t ag 2 t and s ag1 t ag1 t (13)

rocking components may be expressed as


0.5
sin 0 p 2sin 2 0 s 1  sin 0 s
2 2
.. . .
xx t ag 2 t ag 1 t (14)
 cT 1 2sin 2 0 s
1 2
cL

.
Further, it is advised to filter the ag 2 t through a band pass filter with a set of corner frequencies f c
and f max , where f c is the cut-off frequency described above (separating low and high frequency range)
.
and f max is the Nyquist frequency; and filter the ag1 t through a low pass filter with a corner frequency
.
same as f c . This is as per the observation illustrated in context with the auto-spectral densities of ag 1


. if
and ag 2 . Denoting the filter by , and absorbing a negative sign without losing generality, Eq (14)
can be expressed as (dropping off the equivalence sign)
0.5
sin 0 p 2sin 2 0 s 1  sin 0 s
if 2 2 if
.. . .
xx t ag 2 t ag 1 t (15)
 cT 1 2sin 2 0 s
1 2
cL

Denoting two constants as follows


cL cT
C2 and C1 = (16)
sin 0 p 0.5
2sin 0 s 1  sin 0 s
2 2 2

 1 2sin 2 0 s

9
Eq (15) may be rewritten for the rocking component as
if if
.. 1 . 1 .
xx t1 2
ag 2 t ag 1 t
C2 C1 C2
if if
and  (17)
1 1 . . C1
ag 2 t ag 1 t
C1 

Main objective of this paper is to find an appropriate translational motion derivative which can fully
describe the rocking motion in principal plane. Next, an appropriate translational component is defined
such that its derivative takes the form as follows:

1
a ag 2  ag1 (18)

The objective here is to explore the possible linearity between the rocking component and a . Hence,
the parameter  has significant influence on defining the appropriate translational component for
rocking motion.
Calculation of alpha

This parameter (  ) will, depend on event and station. Hence it is recommended to choose  , such
that response spectrum obtained by the appropriate translational motion derivative closely resembles
the response spectrum of that of rocking.
The response spectra of the ground motion is first smoothed with a 20 point hamming window to get
smoothed response spectrum. Distance correlation between rocking and a for square of deviation
spectra is then calculated with varying alpha from 0.1 to 20. Similarly distance correlation between
rocking and a for smoothed spectra is calculated with varying alpha. Correlation of square of deviation
spectra has been plotted against correlation of smoothed spectra in Figure 4a for few stations. A
parameter called total correlation is defined as square root of sum of squares of both correlations. α
which will give maximum total correlation for both deviation spectra and smoothed spectra is chosen
as appropriate alpha. Variation of this total correlation with respect to alpha is shown in Figure 4b for
Event 2. It can be concluded from Figure 4b that even though appropriate alpha is different stations, the
plot of correlation against alpha is flat at maximum correlation for a particular station. Hence without
significant loss of correlation a constant alpha can be assumed.
Characterization Based on Response Spectra

After choosing appropriate alphas, PGAs of rocking motion are compared with that of a in Figure 5.
Some degree of correlation is noted in each of the three events. Table 3 shows the correlation
coefficients obtained between the smoothed spectra of torsional motion vs a g 3 and rocking vs a .
Correlation coefficients obtained are very close to one. Therefore, the component a , defined per Eq
(18), may be considered as the appropriate translational component for rocking acceleration. Similar
conclusion can be made for torsional motion and a g 3 .

These observations will be useful in developing a simplified method of extracting rotational time series
without P-S wave decomposition of the recorded translational ground motion and will be reported
elsewhere separately.

Spatial Variability

Spatial variability of ground motion over the footprint of array may be assessed through variation of
distance correlation (for smoothed spectra) with the distance between stations. Variation between the
mean, mean-plus-sigma and mean-minus-sigma spectra provide a measure of spatial variability under

10
statistical significance. Figure 6, for example, illustrates such a comparison for rocking ground motion
during Event-2 and Figure 7 shows the spatial variation of smoothed spectra (20 points). Similar
variation for the a g 1 is presented in Figures 8 and 9. Remaining figures were not presented here for
brevity.

Both torsional and rocking spectra exhibit significant spatial variability, especially, at the short period
range. Unlike rotational spectra, translational ground motions do not show significant spatial variability.
Further, the spatial variability beyond 1 sec is negligible regardless of the translational components.
These observations are true for all the three events considered here. Hence all the smoothed response
spectra were considered till 1.5 seconds time period and distance correlation between the all the possible
station pairs was calculated. Distance range is divided into bins of space 10m and the correlations falling
in a specific bin were averaged. Obtained correlation ranges (maximum and minimum) were tabulated
in Table 4 for all the events. Relative comparison of spatial variability in three translational components
considered here varies from one event to another. This applies to the torsional and rocking spectra, as
well.

Characterization Based on Auto Spectral Density

Another way of ground motion characterization is through its auto-spectral density (ASD) function that,
in principle, provides the information about frequency content and distribution of energy over a
frequency band. Assuming ergodicity and the recorded time series ( x(t ) ) as a product of finite duration
strong motion window and the sample function of an actual stationary random process, the ASD
function may be calculated using finite FT ( X ( f ) )as follows:

1 *
S xx  f   X  f ,T  X  f ,T  (19)
T
ASD given by Eq (19) is generally sporadic in nature and thereby rendering challenges while extracting
any meaningful conclusion. Often spectral density functions calculated for the recorded ground motions
are smoothened in frequency domain using a weighted window in order to extract meaningful
information including frequency content. Hamming window is used to carry out smoothening in this
paper.

Five hamming points have been considered on each side of the data point in this paper. Smoothened
ASDs are computed for three translational components, namely, a g 1 , a g 2 and a g 3 , and two rotational
components, namely, torsional and rocking on principal plane. Results are available at all the working
surface stations during all the three events. As an example, Figure 10 compares the ASD of a g 1 , without
and with smoothening, at station FA1_1 during Event-1.

Parametric Representation of ASD

The smoothened ASD itself provides much insight into the dominant properties of ground motion.
Bachmann, H et al (1987) estimated the ASD due to forces by walking pedestrians on the bridge. And
its ASD is fitted through normal distribution and log normal distribution to estimate the response of the
bridge. Similarly, the shape of smoothened ASD [Figure 10(b)] has been observed to be resembling a
log normal curve with a dominant frequency and decaying exponentially farther away from that
dominant frequency. Hence, the ASD is characterised here using the log normal form involving three
parameters, namely, i) Amplitude ( A ), ii) Mean frequency ( f m ) and iii) Frequency spread ( f s ), as
follows:

A  (log( f )  log f ) 2 
S ( f )  exp     A * P( f ) (20)
f  2 log
2
 f 

11
where  log f and  log f are the mean and standard deviation of log( f ) , respectively.

Numerically calculated ASD has been approximated through Eq (20) and the required curve-fitting is
not straightforward as there are three variables convoluted in a nonlinear way. Hence mean frequency
and frequency spread are assumed in priori in normal scale and then converted to logarithmic scale as
shown below.

 
 
 ,  log f  log 1  f s 2 
fm
log f  log  (21)
 fs   fm 
 1 f 2 
 m 

Distance correlation is then calculated between the obtained P ( f ) and actual S ( f ) . Now the
procedure is repeated considering a range of frequencies for mean frequency and frequency spread. The
combination of mean frequency and frequency spread, which will give maximum correlation will be
taken as actual mean frequency and frequency spread. Amplitude of ASD is then calculated by linear
regression analysis.

1
A   PT P  PT S (22)

Smoothened ASD is compared with its parametric representation in Figure 11a for one translational
motion. Such comparison is also carried out for the remaining stations along a g 1 , a g 2 and a g 3 , but not
included here for brevity. Close resemblance is observed in most cases.

Note that similar parametric representation [Eq (20)] is not valid for the derivative of translational
motion (which is also closely related to the rotational motion, if selected appropriately). Consider, for
example, a g 3 and denoting X (.) and S (.) , as the FT and ASD, respectively, one may write:

X ag 3 ( f )  i 2 fX ag 3 ( f ) and Sag 3 ag 3 ( f )  (2 f )2 Sag 3 ag 3 ( f ) (23)

Since parametric representation of ASD of a g 3 is given by a log normal form, that of its time derivative
cannot follow the same form owing to Eq (23). However, ASD of any translational motion derivative
may first divided by a factor 4 2 f 2 at the respective frequencies before developing its parametric
representation as follows:

Saa ( f ) A  (log( f )   ) 2 
 exp  
4 2 f 2 f  2 2 
(24)
A  (log( f )   ) 2  
Saa ( f )  4 f  exp  
2 2

 f  2 2 

Finally, Eq (24) also holds good for the rotational motion under the assumption of its linearity with
appropriate translational component and proportionality constant is the reciprocal of apparent wave
velocity (a parameter to be discussed later at length in this paper).

12
Parametric representation is compared with the numerically calculated smoothened ASD of rotational
motion and the comparison is presented here for the selected stations in Event-1 (Figure 11b). Once
again, close resemblance is evident.

Comparison of Parametric Representation of ASD: Rotational and derivative of Appropriate


Translational Components

Figure 12a compares the amplitudes of ASD for rocking acceleration and a in all three events. Similar
data are presented in Figure 12b for torsional acceleration and a g 3 . Results indicate a strong correlation
between rotational motion characteristics and that of proposed appropriate translational motion
derivatives. Note that only correlation of amplitudes is illustrated here and not their equivalence.
Equivalence of these two amplitudes involve a parameter like apparent wave velocity (Note that the
slope of the linear relationship between amplitudes of rotational motion and derivative appropriate
translational motion is not equal to unity). To measure the similarity of shapes, distance correlation has
been calculated between the rotational motion and derivative of appropriate translational motion and it
is shown in Table 5. It can be concluded from Table 5 that there exists a strong correlation between
rotational motion and appropriate translational motion derivative.

Only parameters for the Event 2 have been tabulated in Table 6 for brevity. Table 6 enables similar
comparison for the centre frequency and frequency spread. Unlike correlation of amplitudes,
equivalence of these parameters (centre frequency and frequency spread) in rotational (rocking and
torsional) and derivative of appropriate translational motions are considered in this comparison. The
equivalence is well exhibited by the torsional acceleration and a g 3 . However, the equivalence is not
descent in rocking acceleration and a . This may be attributed to the parameter  , and refinement of
which with due technical basis will improve the equivalence. Note that the objective here is to illustrate
the existence of an appropriate translational component associated with a rotational component, which
is nevertheless evident from the discussion presented above.

Spatial Variability of ASD characteristics

ASD (smoothened and averaged over the array) of the rotational and translational ground motions are
presented in Figures 13 and 14 for the Event-1. Also included are the mean-plus-sigma and mean-
minus-sigma ASD to enable an assessment of the spatial variability. Similar comparison for other two
events are not shown here for brevity. Similar to the procedure followed in spatial variability of response
spectra, spatial variability of ASD characteristics has been studied. ASD obtained through log-normal
approximation has been chosen as a variable to obtain the correlation. Maximum and minimum values
of correlation were shown in Table 7. From Table 7 it can be concluded that both rotational motions
exhibit significant spatial variability in ASD with rocking motion on relatively higher side. The extent
of spatial variability in rotational motion is much higher in Event-3 as compared to the other two events.
Epicentral distance being much smaller (20 km), point source assumption may not be appropriate
leading to a large spatial variability. Spatial variability of translational components is relatively less as
compared to the rotational components, an observation also noted in response spectrum
characterization.

Rotational Invariance of Geometric Mean of ASD (Parametric Representation)

From Tables 6 and 7 it can be told that ASD parameters along a particular direction, by and large,
remain constant over the array for a particular event (for example, Event-2 here). These parameters
differ if another direction is chosen for the same event. In order to explore the rotational invariance of
geometric mean of ASD of two orthogonal horizontal ground motions, a g 1 and a g 3 have been rotated
though an angle form 0o to 90o at the increments of 5o. Resulting ground motion pair for each rotation
angle are denoted as H1 and H2. Parametric representation of ASD for H1-H2 pair is tabulated in Table
8 for some selected angles of rotation for the Event-1 (at station FA1_1). The characteristics obtained

13
are significantly different in two orthogonal directions. Geometric mean of the H1-H2 pair is then
calculated and the variation of ASD parameters is presented in Figure 15. The frequency characteristics
(centre frequency and frequency spread) of the Geomean ASD are found to be nearly constant for a
particular event: regardless of the stations and rotation angles. But the amplitudes exhibit (a nominal)
sinusoidal variation with respect to the rotation angle, which may be explained through simple
trigonometry as follows:

Let ax and a y be the two horizontal acceleration time series (say, recorded) before the rotation. When
these recorded data are rotated through an angle  , the resulting time series are given by

ax '  ax cos   a y sin 


(25)
a y '  -ax sin   a y cos 

Denoting A  f  as the FT of time series a  t  , Eq (25) leads to

Ax ' ( f )  Ax ( f ) cos   Ay ( f )sin 


(26)
Ay ' ( f )  - Ax ( f )sin   Ay ( f ) cos 

Resulting ASDs after rotation may be expressed as

S x ' x '  S xx cos2   S yy sin 2   A sin 2


where A  Re  S xy  (27)
S y ' y '  S xx sin 2   S yy cos2  - A sin 2

Subsequently, the geometric mean is given by

 B  AS AS yy   B
S  S x ' x ' S y ' y '   S yy S xx    sin 4  - xx    cos 4  - 
 2  2 2   2
(28)
2
( S xx - S yy )
where A  Re( S xy ) and B  - A2
4

Eq (28) confirms the sinusoidal variation of amplitude of geomean ASD with respect to the angle of
rotation. Note that the same is not true for the frequency parameters.
Spatial and Event-to-Event Variability of Area of ASD

Product of amplitude and frequency spread of the geomean ASD, which is approximately equal to the
associated area (an approximate measure of the signal energy), has been compared over the footprint of
array in all three events (Table 9). Note the product is fairly uniform over the footprint of array:
coefficient of variation (COV) is 11% for the Event-1 while 6% for the Event-2 and 9% for the Event-
3. Based on the rotational invariance of geomean ASD, any two horizontal directions, including the
directions of recording may be chosen for computing the geomean ASD.
Even though product of amplitude and frequency spread is assumed to be approximately equal to the
area under geomean ASD, spatial variability of actual area is presented in Table 10 for Event 1. A trend
of fairly invariance over the footprint of array is noted for the all the events. Note that geomean ASD is
computed based on a g 1 and a g 3 . Table 10 also includes the area of ASD for three individual
translational components, which shows considerable variability over the foot print of array, even during
a particular event. Similar conclusions to Product of amplitude and frequency spread of the geomean
ASD, can be drawn from here.
Apparent Wave Velocity

14
Assuming non-dispersive plane wave propagation, spatial derivative of resulting displacement field at
any point due to a single type wave incidence is related to its temporal derivative through a measure of
velocity. Bouchon and Aki (1982) defined the ratio of temporal to spatial derivative at any point on the
displacement filed as the apparent wave velocity. Hence, time series of spatial and temporal derivatives
can be obtained from one another through scaling up or down by apparent wave velocity for a non-
dispersive (frequency independent wave velocity) plane wave propagation. Single station procedure
developed by Basu et al. (2012a), as used in this paper, defines torsional motion by the spatial derivative
of SH wave field. Similarly, rocking motion on the principal plane is defined as the spatial derivative
of the P and SV wave fields and followed by the superposition of two components. In line with Bouchon
and Aki (1982), apparent velocities of SH, P and SV waves are defined by Basu et al. (2012b) as follows:

1 ag 3  t  peak agp2  t  agsv2  t 


Csh  , Cp  p peak
, and Csv  sv peak
(29)
2  xy  t   pp  t   pp  t 
peak peak peak

It must be noted that apparent wave velocity is a conceptual idealization and does not exist in practice.
So the above definition of apparent wave velocity has some issues. For example, if apparent velocity of
any wave is calculated using above equation and then computed rotational time series is compared with
the derivative of associated translational time series scaled by the calculated apparent wave velocity,
both the time series exhibit a time delay. If the time delay remains uniform over a considerable time
span, it may be treated as a phase difference and the calculated apparent wave velocity may be assumed
as reasonable (Basu et al., 2012b). On the other hand, apparent wave velocity may not be assumed
reasonable if there is no uniformity in the time delay. Even when the time delay is uniform, the
definition of apparent wave velocity has further problem. For example, apparent wave velocity may
also be defined by using translational acceleration instead of its time derivative and rotational velocity
instead of rotational acceleration. In a similar way, translational velocity and rotation can also be used.
In all three cases, calculated apparent velocities are found to be considerably different. In order to avoid
this problem, calculation of apparent velocity in frequency domain is proposed in this paper. More over
the events considered in this study are close to fault, where contribution of surface wave can be
neglected. Apparent wave velocity of body waves are comparatively less dispersive than the surface
waves. So it can be assumed that apparent wave velocity is constant for all the frequencies.

Assuming that apparent wave velocity for SH wave exists, FT for torsional acceleration may be
expressed as follows:

1 1
 xy  t   ag 3  t    xy  f   ag 3  f  (30)
2Csh 2Csh

Associated ASDs are related through

S ag 3  f 
2
 S xy  f  (31)
4C sh

and leading to the computation of apparent SH- wave velocity as

1 Sag 3  f  peak
Csh  (32)
2 S f
 xy
peak

Similarly, for the P and SV wave, apparent velocities are given by

15
Sa p f Sasv  f 
Cp  , and Csv 
g2 g2
peak peak
(33)
S p  f  S sv  f 
pp pp
peak peak

Here, S a p
g2
 f  and Sa  f  denote the ASD of P and SV contributions to the vertical ground motion;
sv
g2

and S p
pp
 f  and S  f  represent the same for rocking acceleration on the principal plane.
sv
pp

Existing definition of apparent velocity relates the absolute maximum of rotational and derivative of
associated translational components. Such a definition can be based on rotation, rotational velocity and
rotational acceleration, and each of which may lead to a drastically different apparent velocity. Note
that Eq (32) and Eq (33) do not depend on whether the derivative of translational acceleration or
translational velocity is used in computing the apparent wave velocity. But note that it is possible only
when the peak values of the ASD of the rotational and translational components occur at the same
frequency which will not be possible in all cases. Hence, it is proposed that smoothed auto-spectral
densities to be used in calculating the apparent wave velocity which will bring the peak frequencies
close if not equal. Apparent wave velocities thus calculated from smoothed ASD are presented in Table
11 for the three seismic events. The comparison of apparent velocities calculated from smoothed ASD
of acceleration time series and velocity time series are shown in Table 12 for Event 1. From Table 12
it can be concluded that apparent P and SV wave velocities calculated from rotational acceleration and
rotational velocity are close.

Note that apparent SH wave velocity remains fairly uniform over the footprint of array: coefficient of
variation (COV) is 5% for Event-1 while 8% for Events-2 and -3. Mean apparent SH wave velocity in
Table 11 shows the event to event variability: Event-1 (187m/sec) and Event-2 (249m/sec) are recorded
at similar epicentral distance, 66km and 75km, respectively, whereas Event-3 (247m/sec) is recorded
at 20km. Apparent P wave velocities show significant spatial variability over the array. Except Event-
1 (5%), considerable COV is noted in Event-2 (30%) and Event-3 (24%). Mean apparent P wave
velocity also varies from one event to another. Observations on apparent SV wave velocity are also
similar. COV is as high as 16% in Even-3 and is minimum in Event -1 (5%). Mean velocity also varies
from one event to another. Overall, Event-1 leads to a minimum variation of apparent. Even though
does not exist theoretically, the apparent wave velocity provides a simple interpretation of rotational
motion from the derivative of the appropriate translational motion.

Energy Time History, Spectra and Strong Motion Duration


Energy measures can also be used as alternative indices to the response quantities such as forces or
displacements and thereby enabling direct inclusion of duration-related seismic damage. The main
advantage of the energy formulation is the replacement of vector quantities, such as displacements,
velocities and accelerations, by the scalar energy quantities. Let one SDOF moves through an increment
of displacement du. Relative energy input by the effective force peff  t   mug (t ) is given by
t t t
E (t)    mu g (t )du    mu g (t )udt  E (t)   u g (t )ut (34)
m 0
0 0

Here, ug is the ground acceleration and u is the relative velocity of SDOF with respect to ground.
Relative energy imparted per unit mass is often used to characterize the structural response and is
generally believed to be an indicator of the damage potential of ground motion. Note that rigid body
movement of the oscillator is not included in the formulation of cumulative relative energy build up [Eq

16
(34)]. Figure 16 presents an illustration for a 5% damped oscillator with natural period 1 sec and
subjected to the ground motion recorded along a g 1 at station FA1_1 during Event-1.

Note that Eq (34) is specific to the oscillator chosen and hence, to enable a comparative description
over a band of natural periods, energy spectrum is often used: A plot of maximum cumulative relative
energy imparted to a spectrum of SDOFs with constant damping ratio. Examples of such (5% damped)
energy spectra are shown in Figure 17 for a g 1 at FA1_1 in Event-1.

Strong motion duration, often extracted from the recorded accelerograph, is another parameter generally
believed to be correlated with the damage potential. A motion of short duration may not produce enough
load reversals for damaging response to build-up in a structure, even if the amplitude of motion is high.
On the other hand, a motion with moderate amplitude but long duration can cause substantial damage.
The duration of a strong ground motion is related to the time required for the release of accumulated
strain energy by rupture along the fault. As the length, or area, of the fault rupture increases, the time
required for rupture also increases. As a result, strong motion duration increases with increasing
earthquake magnitude and hence, is of interest to the engineers rather than the duration of entire time
history. Amongst various other measures, time laps between two different percentages of peak
cumulative relative energy build up (say, 5%-95%; 10%-90%, etc.) is often considered as the strong
motion duration. Figure 16 presents an illustration based on 2.5%-97.5% of peak cumulative relative
energy build up: T2.5 and T97.5 denote the time required to attain 2.5% and 97.5% of the peak,
respectively; and strong motion duration is given by T95 = T97.5 – T2.5.

Figure 16 presents the case for 1-sec oscillator and the resulting strong motion duration is expected to
be different if any other natural period is selected. Resulting variation of the strong motion duration for
a spectrum of oscillator is presented in Figure 18. While T2.5 remains fairly constant except at the short
period range, T97.5 exhibits significant period sensitivity and hence, the resulting strong motion duration
(vertical offset between T97.5 and T2.5) varies arbitrarily with respect to the chosen oscillator.

As the strong motion duration exhibits significant period sensitivity, it is not a good parameter for
exploring the correlation between appropriate translational and rotational ground motion. Instead,
cumulative relative energy build-up (in %) over time is considered for comparing the rotational and
derivative of appropriate translational motion, and the sample data are presented in Figures 19 (torsional
and a g 3 ) and 20 (rocking and a ). While Figure 19 illustrates a close proximity, Figure 20 indicates
noticeable difference, which is expected to be minimized provided the parameter  is judiciously
selected. To measure the similarity of shapes, distance correlation has been calculated between the
rotational motion and derivative of appropriate translational motion and it is shown in Table 13. It can
be concluded from Table 13 that there exists a strong correlation between rotational motion and
appropriate translational motion derivative.

Percentage energy build up compared above reflects the behaviour of one oscillator ( T  1sec ). In
order to extend such comparison for a spectrum of oscillators, normalized energy spectra (normalized
with respect to peak spectral ordinate) is chosen in this paper. Figure 21 presents the sample data by
comparing the torsional motion and a g 3 at FA1_1 in Event-1. A close resemblance is evident, as
expected. Similar comparison for rocking and a is included in Figure 22. Resulting comparison is not
that decent, especially, at the shorter periods, but is expected to be improved with judicious selection
of the parameter  .

Further, peak spectral energy (maximum value of energy spectrum) is chosen to explore the correlation
in appropriate translational motion, its time derivative and rotational motion. Peak spectral energy of
rotational motion is plotted against that of the derivative of appropriate translational motion, and sample
illustrations are presented in Figures 23 and 24 for the torsional and rocking motions, respectively, in
Event-1. Reasonable correlation is evident in either case. Other two events also indicate similar trend

17
but not presented here for brevity. Finally, appropriate translational motions do not show similar
correlation with respective rotational motions.

Summary and Conclusions


This paper investigates the possible correlation between the resulting rotational ground motion and the
recorded translational components. Torsional motion is related to the horizontal component normal to
the principal plane whereas, rocking motion is contributed from the horizontal and vertical components
on the principal plane. The investigation shows no correlation between the torsional component and
ground motion rotated normal to the principal plane. Instead, significant correlation is noted with
respect to the velocity time series along the normal direction. Rocking motion on the other hand does
not show correlation with any of the translational components, either recorded or rotated. However, a
translational component is defined in this paper as a combination of horizontal and vertical components
on the principal plane and time derivative of which shows much better correlation with rocking
component. Such a translational component is termed as the appropriate translational component for
rocking motion. Correlation is studied through several representations of ground motion. For example,
through PGA, distance correlation between spectral shapes, parametrization of ASD and energy time
history.
ASD of translational ground motion is represented through a log-normal parametric form involving
three parameters, namely, the amplitude, centre frequency and frequency spread. A slightly different
from but involving the same three parameters has been derived for the ASD of rotational components.
However, parameters of ASD for translational components differs significantly from one direction to
another. Geomean ASD, in parametric form, is seen to be nearly invariant. While the centre frequency
and frequency spread exhibit invariance, a nominal sinusoidal variance is noted for amplitude with
respect to the angle of rotation, which is expected. Product of amplitude and frequency spread of the
geomean ASD, which is approximately equal to the associated area has been compared over the
footprint of array and a nominal spatial variability in noted. Further, studies are required to draw
meaningful conclusion.
Apparent velocity is redefined in this paper, but in frequency domain, which do not depend on whether
the derivative of translational acceleration or translational velocity is used in the computing the apparent
velocity. Use of smoothened ASD is recommended for this purpose. Even though does not exist
theoretically, the apparent wave velocity provides a simple interpretation of rotational motion from the
derivative of the appropriate translational motion.

Cumulative relative energy build up and its spectral representation are also included in the correlation
study. The correlation is investigated through the percentage energy build up and peak spectral energy.
Once again, rotational motion and time derivative of the appropriate translational components are found
to be correlated.
In all cases above, torsional motion exhibits fairly good correlation with the time derivative of
appropriate translational component. However, the correlation is not that decent in case of rocking
motion. With proper selection of the parameter alpha and cut-off frequency, one may expect much better
correlation even for the rocking motion also.
References
1. Aki, K. (1967). Scaling law of seismic spectrum. Journal of Geophysical Research, 72(4), 1217-
1231
2. Bachmann, H., and W. Amman. "Vibrations in Structures Induced by Man and Machines.
Structural Engineering Document 3 e, IABSE, Switzerland, 1987, 176 pp."
3. Basu, D. and Giri, S. (2015). Accidental eccentricity in multistory buildings due to torsional
ground motion, Bulletin of Earthquake Engineering, Springer (DOI :10.1007/s10518-015-9788-0).
4. Basu, D., Constantinou, M. C., & Whittaker, A. S. (2014). An equivalent accidental eccentricity
to account for the effects of torsional ground motion on structures. Engineering Structures, 69,
1-11.

18
5. Basu, D., Whittaker, A. S. and Constantinou, M. C. (2012b). "Characterizing the rotational
components of earthquake ground motion." MCEER-12-0005, Multidisciplinary Center for
Earthquake Engineering Research, SUNY, Buffalo, NY.
6. Basu, D., Whittaker, A. S. and Constantinou, M. C. (2013). " Extracting rotational components
of earthquake ground motion using data recorded at multiple stations." Earthquake Engineering
and Structural Dynamics, 42 (3): 451-468.
7. Basu, D., Whittaker, A. S., & Constantinou, M. C. (2012a). Estimating rotational components
of ground motion using data recorded at a single station. Journal of Engineering Mechanics,
ASCE, 138 (9): 1141-1156.
8. Basu, D., Whittaker, A. S., & Constantinou, M. C. (2015). Characterizing rotational
components of earthquake ground motion using a surface distribution method and response of
sample structures. Engineering Structures, 99, 685-707.
9. Bodin, P., Gomberg, J., Singh, S. K. and Santoyo, M. (1997). "Dynamic deformations of
shallow sediments in the valley of Mexico, Part I: three-dimensional strains and rotations
recorded on a seismic array." Bulletin of Seismological Society of America, 87 (3): 528-539.
10. Bolt, B. A. (1988). Earthquakes. New York, W.H. Freeman.
11. Boore, D. M. (2003). Simulation of ground motion using the stochastic method. Pure and
applied geophysics, 160(3-4), 635-676.
12. Bouchon, M. (1979). "Discrete wavenumber representation of elastic wave fields in three-space
dimensions." Journal of Geophysical Research, 84: 3609-3614.
13. Bouchon, M. (1980). "The motion of the ground during an earthquake. Part 1: the case of a
strike-slip fault." Journal of Geophysical Research, 85: 356-366.
14. Bouchon, M. and Aki, K. (1977). "Discrete wave number representation of seismic-source
wave fields." Bulletin of Seismological Society of America, 67: 259-277.
15. Bycroft, G. N. (1980). Soil—foundation interaction and differential ground
motions. Earthquake Engineering & Structural Dynamics, 8(5), 397-404.
16. Castellani, A. and Boffi, G. (1986). "Rotational components of the surface ground motion
during and earthquake." Earthquake Engineering and Structural Dynamics, 14: 751-767.
17. Castellani, A. and Boffi, G. (1989). "On the rotational components of seismic motion."
Earthquake Engineering and Structural Dynamics, 18: 785-797.
18. de la Llera, J. C., & Chopra, A. K. (1994). Accidental and natural torsion in earthquake response
and design of buildings. Earthquake Engineering Research Center, University of California.
19. EC8.1 (2005). EN 1998-1 Eurocode 8: Design of Structures for Earthquake Resistance. Part 1:
General Rules, Seismic Actions and Rules for Buildings.
20. EC8.6 (2005). EN 1998-6 Eurocode 8: Design of Structures for Earthquake Resistance. Part 6:
Towers, Masts and Chimneys.
21. Falamarz-Sheikhabadi, M. R. (2014). Simplified relations for the application of rotational
components to seismic design codes. Engineering Structures, 59, 141-152.
22. Falamarz-Sheikhabadi, M. R., & Ghafory-Ashtiany, M. (2012). Approximate formulas for
rotational effects in earthquake engineering. Journal of Seismology,16(4), 815-827.
23. Falamarz-Sheikhabadi, M. R., & Ghafory-Ashtiany, M. (2015). Rotational components in
structural loading. Soil Dynamics and Earthquake Engineering,75, 220-233.
24. Ghayamghamian, M. R. and Nouri, G. R. (2007). "On the characteristics of ground motion
rotational components using Chiba dense array data." Earthquake Engineering and Structural
Dynamics, 36: 1407-1429.
25. Gomberg, J. (1997). "Dynamic deformations and M 6.7, Northridge, California earthquake."
Soil Dynamics and Earthquake Engineering, 16: 471-494.
26. Hao, H. (1996). Characteristics of torsional ground motions. Earthquake engineering &
structural dynamics, 25(6), 599-610.
27. Hao, T. Y. (2004). ENERGY OF EARTHQUAKE RESPONSE–RECENT
DEVELOPMENTS. ISET Journal of Earthquake Technology, Paper No. 416, Vol. 39, No. 1-
2, March-June 2002, pp. 21-53.
28. Harichandran, R. S. (1991). Estimating the spatial variation of earthquake ground motion from
dense array recordings. Structural Safety, 10(1), 219-233.

19
29. Hart, G. C., Lew, M., & DiJulio, R. M. (1975). Torsional response of high-rise
buildings. Journal of the Structural Division, 101(2), 397-416.
30. Haskell, N. A. (1953). "The dispersion of surface waves in multilayered media." Bulletin of
Seismological Society of America, 43: 17-34.
31. Kameda, H. (1975). "Evolutionary spectra of seismogram by multifilter." Journal of
Engineering Mechanics, ASCE, 101: 787-801.
32. Laouami, N., & Labbe, P. (2002). Experimental analysis of seismic torsional ground motion
recorded by the LSST‐Lotung array. Earthquake engineering & structural dynamics, 31(12),
2141-2148.
33. Lee, V. W. and Trifunac, M. D. (1985). "Torsional accelerograms." Soil Dynamics and
Earthquake Engineering, 4: 132-139.
34. Lee, V. W. and Trifunac, M. D. (1987). "Rocking strong earthquake accelerations." Soil
Dynamics and Earthquake Engineering, 6: 75-89.
35. Li, H.-N., Sun, L.-Y. and Wang, S.-Y. (2002). "Frequency dispersion characteristics of phase
velocities in surface wave for rotational components of seismic motion." Journal of Sound and
Vibration, 258 (5): 815-827.
36. Li, H.-N., Sun, L.-Y. and Wang, S.-Y. (2004). "Improved approach for obtaining rotational
components of seismic motion." Nuclear Engineering and Design, 232: 131-137.
37. Lomnitz, C. (1997). "Frequency response of a strainmeter." Bulletin of Seismological Society
of America, 87 (4): 1078-1080.
38. Makris, N., & Black, C. J. (2004). Evaluation of peak ground velocity as a “good” intensity
measure for near-source ground motions. Journal of Engineering Mechanics, 130(9), 1032-
1044.
39. Menke, W. (1984). Geophysical Data Analysis: Discrete Inverse Theory. International
Geophysics Series, Academic Press, Inc. 45.
40. Newmark, N. M. (1969, January). Torsion in symmetrical buildings. PROCEEDING OF
WORLD CONFERENCE ON EARTHQUAKE ENGINEERING.
41. Niazi, M. (1986). "Inferred displacements, velocities and rotations of a long rigid foundation
located at El Centro differential array site during the 1979 Imperial Valley, California,
earthquake." Earthquake Engineering and Structural Dynamics, 14: 531-542.
42. Penzien, J., & Watabe, M. (1974). Characteristics of 3‐dimensional earthquake ground
motions. Earthquake engineering & structural dynamics, 3(4), 365-373.
43. Politopoulos, I. (2010). Response of seismically isolated structures to rocking‐type
excitations. Earthquake Engineering & Structural Dynamics, 39(3), 325-342.
44. Rathje, E. M., Abrahamson, N. A., & Bray, J. D. (1998). Simplified frequency content estimates
of earthquake ground motions. Journal of Geotechnical and Geoenvironmental Engineering.
45. Singh, S. K., Santoyo, M., Bodin, P. and Gomberg, J. (1997). "Dynamic deformations of
shallow sediments in the valley of Mexico, Part II: single-station estimates." Bulletin of
Seismological Society of America, 87 (3): 540-550.
46. Spudich, P., Steck, L. K., Hellweg, M., Fletcher, J. B. and Baker, L. (1995). "Transient stress
at Parkfield, California, produced by the M 7.4 Landers earthquake of June 28, 1992:
Observations from the UPSAR dense seismograph array." Journal of Geophysical Research,
100: 675-690.
47. Sucuoglu, H. A. L. U. K., & Nurtug, A. (1995). Earthquake ground motion characteristics and
seismic energy dissipation. Earthquake Engineering and Structural Dynamics
48. Sugito, M., Goto, H. and Aikawa, F. (1984). "Simplified separation technique of body and
surface waves in strong motion accelerograms." Structural Engineering/Earthquake
Engineering, Proc. Japan Society of Civil Engineers: 71-76.
49. Trifunac, M. D. (1982). "A note on rotational components of earthquake motions for incident
body waves." Soil Dynamics and Earthquake Engineering, 1: 11-19.
50. Tso, W. K., Zhu, T. J., & Heidebrecht, A. C. (1992). Engineering implication of ground motion
A/V ratio. Soil Dynamics and Earthquake Engineering, 11(3), 133-144.
51. Wolf, J. P., Obernhueber, P. and Weber, B. (1983). "Response of a nuclear plant on aseismic
bearings to horizontally propagating waves." Earthquake Engineering and Structural Dynamics,
11: 483-499.

20
52. Wong, H. L. and Trifunac, M. D. (1979). "Generation of artificial strong motion
accelerograms." Earthquake Engineering and Structural Dynamics, 7: 509-527.
53. Zembaty, Z. (2009a). "Rotational seismic code definition in Eurocode 8, Part 6, for slender
tower-shaped structures." Bulletin of Seismological Society of America, 99 (2B): 1483-1485.
54. Zembaty, Z. (2009b). "Tutorial on surface rotation from wave passage effects: Stochastic
spectral approach." Bulletin of Seismological Society of America, 99 (2B): 1040-1049.
55. Zembaty, Z. and Boffi, G. (1994). "Effect of rotational seismic ground motion on dynamic
response of slender towers." European Earthquake Engineering, 8: 3-11.
56. Zerva, A., & Zervas, V. (2002). Spatial variation of seismic ground motions: An
overview. Applied Mechanics Reviews, 55(3), 271-297.

21
Figure 1: Location of free surface stations

3.5 0.07
Spectral Acceleration (m/sec2)

3 0.06
Spectral Acceleration

2.5 0.05
(rad/sec2)

2 0.04
1.5 0.03
PGA 0.02
1
0.5 0.01
0 0
0 1 2 3 0 1 2 3
Time Period (sec) Time Period (sec)
(a) Response spectrum for ag1 (b) Response spectrum for Torsion
Figure 2: Spectral Acceleration plot for Event 1 at Station FA1_1

Event 1 Event 2 Event 3


32
PGA of ȧg3 (m/sec3)

27 R² = 0.9698

22
17 R² = 0.9788

12
R² = 0.9757
7
0.02 0.04 0.06 0.08
PGA of Torsion (rad/sec2)
Figure 3: PGA comparison of Spectra of torsion and ȧg3

22
1 1.4
Correlation of squate of deviation
FA1_2 FA1_1
0.9 1.2
0.8 FA1_3 FA1_2

Total correlation
0.7 FA1_4 1 FA1_3
0.6 FA1_5 0.8
FA1_4
spectra

0.5 FA1_5
FA2_2 0.6
0.4 FA2_1
0.3 FA2_1
0.4 FA2_2
0.2 FA2_3
FA2_3
0.1 0.2
FA2_4
FA2_4
0 0
0 0.2 0.4 0.6 0.8 1 FA2_5
0 4 8 12 16 20
Correlation of smoothed spectra FA3_5
alpha
(a) Distance correlation of square of deviation (b) Distance correlation of smoothed spectra
spectra against alpha against alpha
Figure 5: Variation of correlation coefficient against alpha for Event 2

Event 1 Event 2 Event 3


20
R² = 0.8271
PGA of ȧ (m/sec3)

15

10 R² = 0.6096

5 R² = 0.4514

0
0
0.02 0.04 0.06 0.08 0.1
PGA of Rocking (rad/sec )2

Figure 5: PGA comparison of Spectra of rocking and ȧ

0.4
FA1_1
0.35 FA1_2
Spectral Acceleration (rad/sec2)

FA1_3
0.3 FA1_4
FA1_5
0.25 FA2_1
FA2_2
0.2 FA2_3
FA2_4
0.15 FA2_5
FA3_5
0.1 Average
Avg+Std
0.05 Avg-Std

0
0 0.5 1 1.5 2 2.5 3
Time Period (seconds)
Figure 6: Spatial variability of rocking spectra - (Event-2)

23
0.3
FA1_1

Smoothed Spectral Acceleration


FA1_2
0.25
FA1_3
FA1_4
0.2
FA1_5
(rad/sec2)
FA2_1
0.15
FA2_2
FA2_3
0.1
FA2_4
FA2_5
0.05
FA3_5

0
0 0.5 1 1.5 2 2.5 3
Time Period (seconds)
Figure 7: Spatial variability of smoothed rocking spectra - (Event-2)

0.8 FA1_1
FA1_2
0.7 FA1_3
Spectral Acceleration (Sa/g)

FA1_4
0.6 FA1_5
FA2_1
0.5 FA2_2
FA2_3
0.4 FA2_4
FA2_5
0.3 FA3_5
Average
0.2

0.1

0
0 0.5 1 1.5 2 2.5 3
Time Period (seconds)
Figure 8: Spatial variability of ag1 spectra - (Event-2)
0.7 FA1_1
FA1_2
Smiithed Spectral Acceleration

0.6 FA1_3
FA1_4
0.5 FA1_5
FA2_1
0.4 FA2_2
(Sa/g)

FA2_3
0.3 FA2_4
FA2_5
FA3_5
0.2

0.1

0
0 0.5 1 1.5 2 2.5 3
Time Period (seconds)
Figure 9: Spatial variability of smoothed ag1 spectra - (Event-2)

24
0.09 0.04

Auto Spectral Density


Auto Spectral Density

0.08 0.035
0.07 0.03

(m2/sec4/Hz)
(m2/sec4/Hz)

0.06 0.025
0.05 0.02
0.04
0.015
0.03
0.02 0.01
0.01 0.005
0 0
0 1 2 3 0 1 2 3
Frequency (Hz) Frequency (Hz)
(a) Unsmoothed spectral density (b) Smoothened spectral density
Figure 10: Auto-Spectral Density of ag1 at FA1_1 for Event 1

Actual log normal Actual log normal

0.06 8E-5
Auto Spectral Density

Auto Spectral Density

0.05 7E-5
6E-5
(rad2/sec4/Hz)
(m2/sec4/Hz)

0.04 5E-5
0.03 4E-5
0.02 3E-5
2E-5
0.01 1E-5
0 0E+0
0 2 4 6 8 10 0 2 4 6 8 10
Frequency (Hz) Frequency (Hz)

(a) Auto-Spectral Density of ag1 at FA1_1 (b) Auto-Spectral Density of Rocking


for Event 3 motion at FA1_2 for Event 1
Figure 11: Parametrisation of Auto-Spectral Density

Event 1 Event 2 Event 3 Event 1 Event 2 Event 3


0.04 0.1
Amplitude of ȧ g3

R² = 0.9015
Amplitude of ȧ

R² = 0.7201
(m2/sec4/Hz3)

(m2/sec4/Hz3)

0.03
0.02 R² = 0.6442 0.05
0.01 R² = 0.5192
R² = 0.4631
R² = 0.8722
0 0
0.0E+00 1.5E-06 3.0E-06 0.0E+00 1.0E-07 2.0E-07 3.0E-07
Amplitude of rocking (rad2/sec4/Hz3) Amplitude of torsion (rad2/sec4/Hz3)
(a) Rocking motion against ȧ (b) Torsional motion against ȧg3
Figure 12: Comparison of amplitudes of rotational motion and appropriate translational motion
derivatives

25
2.0E-5
Auto Spectral Density (rad2/sec4/Hz) Average Avg+Std Avg-Std

1.5E-5

1.0E-5

5.0E-6

0.0E+0
0 1 2 3 4 5 6 7 8

-5.0E-6
Frequency (Hz)

Figure 13: Spatial variability of torsional motion - (Event-1)

0.1
Auto Spectral Density (m2/sec4/Hz)

0.08 Average Avg+Std Avg-Std

0.06

0.04

0.02

0
0 1 2 3 4 5 6 7 8
-0.02
Frequency (Hz)

Figure 14: Spatial variability of ag3 motion - (Event-1)

26
0.7 FA1_1 1.4 FA1_1 0.8 FA1_1
Amplitude of ASD (m/sec2/Hz)

0.6 FA1_2 1.2 FA1_2 0.7 FA1_2

Frequency spread (Hz)


Centre frequency (Hz)
FA1_3 FA1_3 0.6 FA1_3
0.5 1
FA1_4 FA1_4 0.5 FA1_4
0.4 FA1_5 0.8 FA1_5 FA1_5
0.4
0.3 FA2_1 0.6 FA2_1 FA2_1
0.3
FA2_2 FA2_2 FA2_2
0.2 0.4
FA2_3 FA2_3 0.2 FA2_3
0.1 FA2_5 0.2 FA2_5 0.1 FA2_5
0 FA3_1 0 FA3_1 0 FA3_1
0 15 30 45 60 75 90 FA3_2 0 15 30 45 60 75 90 FA3_2 0 15 30 45 60 75 90 FA3_2
Angle (Degrees) Angle (Degrees) Angle (Degrees)

(a) Variation of amplitude with rotation (b) Variation of centre frequency with rotation (c) Variation of frequency spread with rotation
Figure 15: Variation ASD characteristics with rotation for Event 1

27
0.30 0.35

Spectral Energy (m2/sec2)


Energy time history
0.25 0.30
0.20 0.25
(m2/sec2)

0.15 0.20
T2.5 T97.5
0.10 0.15
0.05 0.10
0.05
0.00
0 10 20 30 T95 40 0.00
0 1 2 3 4 5
Time (Seconds)
Time Period (seconds)
Figure 16: Relative energy time history for Event Figure 17: Relative energy spectra for Event 1 at
1 at FA1_1 for ag1 for T = 1 second FA1_1 for a g 1

T_2.5 T_97.5
40.0
Time (seconds)

30.0

20.0

10.0

0.0
0 1 2 3 4 5
Time Period (seconds)
Figure 18: Variation of T2.5 and T97.5 for Event – 1 at FA1_1 for a g 1

Torsional Motion Derivative of ag3 Rocking Motion Derivative of a


Energy time history (%)
Energy time history (%)

100 100
80 80
60 60
40 40
20 20
0 0
0 10 20 30 40 0 10 20 30 40
Time (Seconds) Time (Seconds)

Figure 19: Comparison of energy build-up for Figure 20: Comparison of energy build-up for
torsional motion and a g 3 for Event 1 at FA1_1 rocking motion and a for Event 1 at FA1_1 for
for T = 1 second T = 1 second

28
Torsional Motion Derivative of ag3 Rocking Motion Derivative of a
1.0
Normalized Spectral

Normalized Spectral
1.0
0.8 0.8

Energy
Energy

0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1 2 3 0 4 5 1 2 3 0 4 5
Time Period (seconds) Time Period (seconds)
Figure 21: Comparison of normalized spectral Figure 22: Comparison of normalized spectral
energy for torsional motion and a g 3 for Event 1 energy for rocking motion and a for Event 1 at
at FA1_1 FA1_1

0.035 Peak spectral energy of 0.15


0.03 R² = 0.3729 R² = 0.4129
Peak spectral energy of

0.13
rocking (m2/sec2)
0.025
torsion (m2/sec2)

0.02 0.11
0.015 0.09
0.01
0.07
0.005
0 0.05
3200 3400 3600 3800 4000 2500 2900 3300 3700 4100
Peak energy of ȧg3 Peak energy of ȧ

Figure 23: Comparison of peak energies of a g 3 Figure 24: Comparison of peak energies of a
and torsional motion for Event-1 and rocking motion for Event-1

29
Table 1: Strong motion events considered
Sl
No. Description Event-1 Event-2 Event-3
1. Date May 20, 1986 November 14, 1986 January 16, 1986
24deg-45min-
2. Latitude 24deg-4min-54sec 23deg-59min-30.5sec 46.2sec
121deg-35min- 121deg-49min-59.4 121deg-57min-40.1
3. Longitude 29.4 sec sec sec
4. Focal Depth (kM) 15.8 15 10.2
5. Local Magnitude 6.2 6.5 6.1
6. Epicentral Distance (kM) 66 75 20

Table 2: Spectral characteristics of translational motion


Torsional Rocking in
ag 2 a g1 ag 3 motion principal plane ag1 ag 2 ag 3
pga pga pga pga pga pga pga pga
Station (g) (g) (g) (rad/sec2) (rad/sec2) (rad/sec3) (rad/sec3) (rad/sec3)
FA1_1 0.05 0.15 0.15 0.03 0.026 37.41 13.65 12.68
FA1_2 0.04 0.18 0.14 0.06 0.034 33.25 13.50 33.46
FA1_3 0.04 0.23 0.15 0.04 0.046 39.65 13.58 23.19
FA1_4 0.04 0.20 0.15 0.05 0.042 38.94 10.37 24.22
FA1_5 0.04 0.21 0.15 0.06 0.035 44.02 8.96 28.33
FA2_1 0.05 0.18 0.15 0.03 0.027 39.07 15.34 24.85
FA2_2 0.07 0.21 0.15 0.05 0.044 38.08 16.77 23.11
FA2_3 0.05 0.21 0.15 0.06 0.035 37.57 14.33 31.42
FA2_5 0.04 0.20 0.15 0.05 0.028 34.20 9.99 23.62
FA3_1 0.06 0.15 0.15 0.04 0.025 40.40 14.93 17.96
FA3_2 0.05 0.16 0.14 0.04 0.027 37.91 13.59 22.23
FA3_3 0.05 0.19 0.15 0.04 0.040 20.70 5.39 20.31
FA3_4 0.04 0.20 0.16 0.06 0.041 33.35 12.23 28.78
FA3_5 0.04 0.21 0.18 0.05 0.030 34.73 15.48 30.23

Table 3: Correlation coefficients – Response spectra


Torsional Motion Rocking Motion
Stations Event 1 Event 2 Event 3 Event 1 Event 2 Event 3
FA1_1 0.990 0.996 0.998 0.92 0.95 0.97
FA1_2 0.991 0.996 0.996 0.92 0.98 0.99
FA1_3 0.990 0.999 0.999 0.95 0.94 0.95
FA1_4 0.992 0.998 0.999 0.95 0.96 0.95
FA1_5 0.992 0.998 0.999 0.96 0.98 0.95
FA2_1 0.995 0.997 0.998 0.87 0.95 0.97
FA2_2 0.993 0.997 0.999 0.91 0.98 0.96
FA2_3 0.988 0.997 0.999 0.94 0.98 0.96
FA2_4 0.996 1.00
FA2_5 0.993 0.997 0.998 0.93 0.99 0.96
FA3_1 0.993 0.999 0.998 0.88 0.98

30
FA3_2 0.990 0.86
FA3_3 0.995 0.998 0.92 0.99
FA3_4 0.995 0.88
FA3_5 0.997 0.997 0.86 0.97

Table 4: Spatial variation of correlation – Response spectra


ag1 ag2 ag3 Torsional motion Rocking motion
Min 0.888 0.951 0.966 0.946 0.791
Event 1
Max 0.995 0.986 0.998 0.991 0.972
Min 0.901 0.972 0.887 0.950 0.963
Event 2
Max 0.996 0.993 0.988 0.991 0.992
Min 0.981 0.970 0.974 0.965 0.957
Event 3
Max 0.993 0.993 0.993 0.999 0.991

Table 5: Correlation coefficients - ASD


Torsional Motion Rocking Motion
Stations Event 1 Event 2 Event 3 Event 1 Event 2 Event 3
FA1_1 0.983 0.998 0.669 0.935 0.857 0.860
FA1_2 0.983 0.993 0.960 0.983 0.909 0.813
FA1_3 0.983 0.999 0.981 0.954 0.845 0.801
FA1_4 0.981 0.999 0.898 0.995 0.795 0.891
FA1_5 0.981 0.992 0.922 0.982 0.836 0.877
FA2_1 0.993 0.998 1.000 0.940 0.913 0.830
FA2_2 0.983 1.000 0.991 0.878 0.945 0.809
FA2_3 0.981 0.993 0.922 0.978 0.877 0.888
FA2_4 0.993 0.897
FA2_5 0.987 1.000 0.991 0.983 0.918 0.856
FA3_1 0.983 0.898 0.988 0.750
FA3_2 0.983 0.903
FA3_3 0.983 0.791 0.878 0.809
FA3_4 0.995 1.000
FA3_5 1.000 0.993 0.976 0.889

Table 6: Log-normal approximation of ASD for Event -2


Centre frequency (Hz) Frequency spread (Hz)
Torsional Rocking Torsional Rocking
Stations motion ag 3 motion a motion ag 3 motion a
FA1_1 1.39 1.49 1.59 1.20 1.09 1.18 0.89 0.89
FA1_2 1.29 1.49 1.88 1.20 0.99 1.18 1.48 0.89
FA1_3 1.68 1.88 2.17 1.00 1.48 1.77 1.97 0.70
FA1_4 1.68 1.88 1.98 1.10 1.48 1.77 1.38 0.79
FA1_5 1.68 2.17 2.27 1.10 1.48 2.16 1.97 0.79
FA2_1 1.20 1.29 1.68 1.00 0.89 0.99 1.38 0.70
FA2_2 1.10 1.10 1.78 1.29 0.70 0.70 1.28 0.89

31
FA2_3 1.39 1.78 1.88 1.29 1.09 1.67 1.28 0.99
FA2_4 1.49 1.98 1.88 1.39 1.18 1.97 1.18 0.99
FA2_5 1.68 1.68 2.08 1.10 1.48 1.48 1.97 0.79
FA3_5 2.27 3.15 1.98 1.50 2.55 3.02 0.31 0.60

Table 7: Spatial variation of correlation – ASD


ag1 ag2 ag3 Torsional motion Rocking motion
Min 0.970 0.912 0.995 0.978 0.912
Event 1
Max 0.998 0.999 0.999 0.993 0.975
Min 0.978 0.771 0.970 0.980 0.626
Event 2
Max 0.998 0.982 0.995 0.996 0.983
Min 0.921 0.694 0.711 0.744 0.533
Event 3
Max 0.983 0.992 0.881 0.953 0.864

Table 8: Variation ASD characteristics with rotation for Event - 1


H1 H2
Angle A fc  A fc 
(Degrees) (g) (Hz) (Hz) (g) (Hz) (Hz)
0 0.0034 1.12 0.51 0.0040 1.03 0.51
15 0.0029 1.22 0.61 0.0045 1.03 0.51
30 0.0025 1.22 0.71 0.0048 0.93 0.42
45 0.0026 1.22 0.71 0.0046 1.03 0.51
60 0.0029 1.12 0.61 0.0042 1.03 0.51
75 0.0034 1.12 0.61 0.0036 1.12 0.61
90 0.0040 1.03 0.51 0.0034 1.12 0.51

Table 9: Product of amplitude and frequency spread of Geomean ASD


Event 1 Event 2 Event 3
FA1_1 0.0311 0.0412 0.0275
FA1_2 0.0316 0.0465 0.0318
FA1_3 0.0340 0.0420 0.0349
FA1_4 0.0346 0.0440 0.0347
FA1_5 0.0387 0.0452 0.0330
FA2_1 0.0311 0.0411 0.0281
FA2_2 0.0316 0.0394 0.0285
FA2_3 0.0345 0.0374 0.0335
FA2_4 0.0454
FA2_5 0.0345 0.0392 0.0321
FA3_1 0.0299
FA3_2 0.0296 0.0267
FA3_3 0.0330 0.0301
FA3_4 0.0357
FA3_5 0.0435 0.0401
COV (%) 10.62 6.10 9.12

32
Table 10: Area under geomean ASD – Event 1
ag1 ag2 ag3 GM
FA1_1 0.0375 0.0068 0.0446 0.0392
FA1_2 0.0367 0.0068 0.0480 0.0401
FA1_3 0.0480 0.0064 0.0475 0.0449
FA1_4 0.0466 0.0057 0.0464 0.0437
FA1_5 0.0469 0.0058 0.0477 0.0443
FA2_1 0.0422 0.0080 0.0442 0.0409
FA2_2 0.0454 0.0077 0.0441 0.0421
FA2_3 0.0472 0.0065 0.0538 0.0476
FA2_4 0.0459 0.0060 0.0511 0.0451
FA3_1 0.0404 0.0070 0.0450 0.0406
FA3_2 0.0393 0.0061 0.0451 0.0398
FA3_3 0.0480 0.0048 0.0533 0.0480
FA3_4 0.0454 0.0056 0.0526 0.0465
FA3_5 0.0413 0.0056 0.0498 0.0403
Mean 0.0436 0.0064 0.0481 0.0431

Table 11: Apparent wave velocities (m/sec)


Event 1 Event 2 Event 3
Station11 SH P SV SH P SV SH P SV
FA1_1 201 814 195 234 1976 172 237 1027 149
FA1_2 197 787 192 247 897 203 231 675 157
FA1_3 181 717 188 267 1175 206 261 679 150
FA1_4 189 779 190 230 1220 182 242 694 155
FA1_5 183 764 190 272 923 180 260 717 161
FA2_1 182 720 184 270 1214 266 247 1328 173
FA2_2 202 818 201 237 707 150 275 849 184
FA2_3 195 787 191 233 1146 186 212 758 162
FA2_4 243 1086 186
FA2_5 181 724 181 281 957 172 260 744 166
FA3_1 193 805 194 229 950 152
FA3_2 192 790 195
FA3_3 169 687 160 269 700 192
FA3_4 176 812 191
FA3_5 183 829 196 222 871 182
Mean 187 773 189 249 1107 190 247 829 164
Std Dev 9.7 43.0 9.4 20.1 331.9 29.5 19.4 202.2 14.1
COV 0.05 0.05 0.05 0.08 0.3 0.16 0.08 0.24 0.09

Table 12: Apparent wave velocity (m/sec) (Event 1)


P velocity P velocity SV velocity SV velocity
Stations (acceleration) (velocity) (acceleration) (velocity)
FA1_1 814 813 195 195

33
FA1_2 787 797 192 194
FA1_3 717 739 188 191
FA1_4 779 750 190 182
FA1_5 764 770 190 191
FA2_1 720 743 184 190
FA2_2 818 741 201 192
FA2_3 787 767 191 190
FA2_4 724 821 181 187
FA2_5 805 803 194 193
FA3_1 790 789 195 194
FA3_2 687 716 160 178
FA3_3 812 810 191 190
FA3_4 829 829 196 196
FA3_5 814 813 195 195

Table 13: Correlation coefficients – Energy time history


Torsional Motion Rocking Motion
Stations Event 1 Event 2 Event 3 Event 1 Event 2 Event 3
FA1_1 0.969 0.990 0.784 0.927 0.855 0.849
FA1_2 0.976 0.987 0.945 0.968 0.906 0.803
FA1_3 0.968 0.988 0.980 0.950 0.837 0.791
FA1_4 0.968 0.990 0.887 0.987 0.781 0.882
FA1_5 0.967 0.980 0.907 0.976 0.828 0.871
FA2_1 0.986 0.994 0.988 0.933 0.903 0.823
FA2_2 0.973 0.991 0.979 0.866 0.942 0.800
FA2_3 0.978 0.979 0.907 0.966 0.870 0.885
FA2_4 0.984 0.889
FA2_5 0.978 0.993 0.984 0.980 0.910 0.840
FA3_1 0.976 0.895 0.982 0.741
FA3_2 0.972 0.894
FA3_3 0.980 0.782 0.870 0.804
FA3_4 0.993 0.998
FA3_5 0.990 0.987 0.969 0.885

34

You might also like