Download as pdf or txt
Download as pdf or txt
You are on page 1of 49

Bulletin of Earthquake Engineering

https://doi.org/10.1007/s10518-021-01046-8

ORIGINAL ARTICLE

Natural period and vertical distribution of base shear


in confined masonry buildings using ambient vibration test

Pranav Chakra‑Varthy1 · Dhiman Basu1 

Received: 11 September 2020 / Accepted: 16 January 2021


© The Author(s), under exclusive licence to Springer Nature B.V. part of Springer Nature 2021

Abstract
In semi-urban setting where availability of land can afford construction of low-rise build-
ings, confined masonry may compete with other alternatives of seismic resilient system
provided well-articulated design standards and construction guidelines are available. Most
seismic standards do not make explicit recommendations on the natural period and vertical
distribution of base shear for the design of confined masonry buildings. In such a case, one
of the two alternatives, such as (1) reinforced concrete (RC) frame building with masonry
infill walls and (2) RC frame building with structural walls, is tacitly extrapolated. This
paper is first aimed to explore the possible recommendations from the ambient vibration
testing of a class of confined masonry building stock. Nine (G + 3) confined masonry hos-
tel buildings are considered for Ambient Vibration Testing (AVT). Recorded signatures are
processed and modal characteristics (primarily restricted to the first triplet of fundamental
modes) are extracted. Each building is modelled numerically and fine-tuned followed by a
comparison of natural frequencies and mode shapes in numerical model and experimen-
tal results. The fine-tuned numerical models are analysed against a set of recorded ground
motions. Possible design recommendations for natural period and distribution of base
shear along the height are the key contributions. Empirical equation for natural periods is
derived from the seismic code recommendation on that of reinforced concrete (RC) build-
ings but removing the bias contributed from the height shorter than one storey while using
the experimental results. Distribution of the base shear along the height follows a parabolic
profile with an exponent close to 0.4. Results of AVT indicate the inherent damping ratio
on an average of about 5% which, however, may not be directly used for seismic excitation.
The building stock used for AVT in this paper does not include considerable variations in
height and different varieties of confined masonry constructions. Therefore, recommenda-
tions of this paper should be verified against a larger size of dispersed building stock.

Keywords  Confined masonry building · Natural period · Damping ratio · Base shear
distribution · Ambient vibration test · Merging strategy of sensors

* Dhiman Basu
dbasu@iitgn.ac.in
1
Department of Civil Engineering, Indian Institute of Technology Gandhinagar, Gandhinagar, India

13
Vol.:(0123456789)
Bulletin of Earthquake Engineering

1 Introduction

Confined masonry (CM) constructions have been reported to be explored as an alternate


seismic resilient system to ordinary/reinforced masonry constructions and also even rein-
forced concrete constructions in various parts of the world including the seismically vul-
nerable region of Latin America, Iran, Indonesia, China, Italy, etc. A CM building typically
comprises of masonry walls (made of clay or concrete brick/block masonry units) and rein-
forced concrete (RC) confining elements (such as tie-beams and tie-columns) enclosing the
walls in horizontal and vertical directions, respectively (Meli et al. 2011). The reinforce-
ment is provided in these confining elements only, whereas the masonry walls are usu-
ally not reinforced though horizontal reiforcement is often preferred to enhance the shear
resistance. Lateral resistance is contributed from the confinement provided by horizontal
and vertical RC tie-elements, and bond at the interface between masonry walls and these
confining elements. These two features were perceived in Indian construction practice even
more than hundred years ago. For example, “Assam type housing” successfully witnessed
and withstood the 1897 Assam earthquake without major collapse particularly owing to the
confining timber elements around light weight “ikra walls”. Lesson learnt from 1931 Balu-
chistan earthquake led to an innovative masonry bond and that was adopted in Quetta while
constructing several bungalows for Indian Railways. These bungalows only withstood sub-
sequently the 1935 earthquake that otherwise killed ~ 25,000 persons. However, CM is not
currently practiced in India as these two features have been lost over time. Several factors
contributed to this shift including the lack of available design standards. Available design
standards for RC and steel constructions allow seismic resilient high-rise buildings even
in the areas prone to damaging earthquakes. This is particularly important in urban set-
ting amidst land crisis where CM constructions do not fit. The role of structural design in
the art of building constructions is well articulated by Foraboschi (2016). Nevertheless,
in semi-urban setting where availability of land can afford construction of low-rise build-
ings, CM may compete with other alternatives of seismic resilient system provided well-
articulated design standards and construction guidelines are available. Also, masonry does
not limit itself to only one structural material (Foraboschi 2019). This expectation stems
from the facts that CM buildings withstood the effects of major earthquakes since the last
century without collapse and largely reduced the number of fatalities in these earthquakes
(Basu 2019). Some of these events are 2010 Maule, Chile earthquake (M 8.8), 2007 Pisco,
Peru earthquake (M 8.0), 1995 Manzanillo Earthquake (Mw 8.0), 1985 Llolleo, Chile
earthquake (M 7.8), 1985 Michoacán earthquake (M 8.1), 1979 Guerrero earthquake (M
7.6), 1939 Chillán earthquake (M 7.8), 1908 Messina, Italy earthquake (M 7.2) etc. Never-
theless, CM walls with opening(s) typically experienced significant damages, for example,
during 2010 Chile earthquake, especially when the confinement was inadequate (Astroza
et al. 2012).
Most seismic standards for RC (with/out masonry infill) and steel structures recommend
force based seismic design at the preliminary stage with an option of detailed performance
assessment using displacement-based framework at the final stage. The force based seis-
mic design comprises of the following four steps: (1) Empirical estimate of natural period;
(2) Arriving at the design base shear using regional seismicity/vulnerability and expected
performance of chosen structural configuration; (3) Height wise distribution of design base
shear to construct the lateral load profile for equivalent static analysis; (4) Application of
lateral load profile at an offset from the mass centre (MC) at the respective floor levels in
order to account for the possible effect of torsion (natural as well as accidental). Seismic

13
Bulletin of Earthquake Engineering

force demands thus calculated can be combined with the other load effects (including
appropriate partial safety factors) to ascertain the design member force resultants. Prior
to a detailed displacement-based procedure at the final stage, often seismic displacement
demands from the equivalent static analysis is scaled up by appropriate response reduc-
tion factor (as used in the calculation of design base shear) for the purpose of stability
assessment. In principle, the equivalent static analysis framework also applies to the CM
buildings as recommended by the available guidelines provided the plan layout and overall
configuration conform to the “simplicity”.
While adopting the same force-based framework to CM buildings, several details
need careful deliberation. Table  1 summarizes the recommendations of Indian standard
(IS 1893-Part I 2016) along with some of the international standards for the natural peri-
ods of building structures. Comparison of these different recommendations is beyond the
scope of present paper. However, the takeaway is that the empirical equation of natural
period through experimental studies exclusively on confined masonry building stock is not
yet reported. Regardless of the material properties and structural systems, most seismic
codes including IS 1893-Part I (2016) assume the damping ratio as 5% of critical for all
buildings. Turek et al. (2006) reported the damping ratio as less than 5% (2.5–4.1% at first
mode) for masonry buildings in Canada using ambient vibration test (AVT). Nema and
Basu (2019) reported 5% damping for confined masonry buildings using AVT on three
buildings recently constructed at the residential campus of Indian Institute of Technology
Gandhinagar. However, much larger building stock is required to be tested to confirm this
expectation. Nevertheless, the damping computed using AVT may not reflect the true sce-
nario during a seismic excitation. Most seismic code recommendations on the vertical dis-
tribution of design base shear are based on the behaviour of RC and steel buildings. CM
buildings are expected to deform predominantly in a shear mode and thereby requiring a
careful assessment on the height wise distribution of design base shear. Similar study is
not yet reported in the prior art. Response reduction factor (while arriving at the design
base shear) to be used with CM buildings may also differ from that of RC buildings.

1.1 Ambient vibration testing: scope, limitation and application

Also known as Operational Modal Analysis (OMA), AVT is an output-only analysis where
the response of structure is only considered. It has several advantages over the Force Vibra-
tion Testing (FVT) including (1) Ambient sources like wind, pedestrians, vehicles etc.
serve the purpose of excitation leading to the feasibility of testing without disturbing the
functionality; and (2) Large force required in FVT (Trifunac 1972) may cause undesirable
damage. However, AVT requires the sensors of high sensitivity and thus making it expen-
sive to have large number of them. Certain merging and referencing techniques reported in
the prior art have been widely used to merge data from different sensors deployed at vari-
ous locations in a structure and at different instants of time (Peeters and De Roeck 1999;
Magalhaes et al. 2012; Brownjohn et al. 2010). This is to overcome the limitation imposed
by number of sensors while maintaining the accuracy of estimating the modal parameters.
AVT was mostly performed on RC and steel buildings, however, there were only few
studies on masonry buildings (Vestroni et al. 1996; De Sortis et al. 2005; Turek et al. 2006;
Nema and Basu 2019). Only one of these studies (Nema and Basu 2019) is related to the
modern CM buildings in India, while the remaining studies are focused on older unrein-
forced masonry buildings in Italy, Canada and Algeria. Therefore, experimental (AVT) esti-
mation of modal characteristics of CM buildings is still at its infancy.

13

Table 1  Empirical equations for natural period according to international seismic codes
Reinforced concrete frames Reinforced concrete and masonry wall Remarks Code/standard (country)
structures

13
3∕4 3∕4 Ae(m2) = the minimum cross-sectional area Uniform Building Code (1997) (USA)
T = Ct hn T = Ct hn
Ct = 0.0731 where in any horizontal plane in the first storey of
a shear wall;
Ct = 0.0488 De (m) = the length of the same shear wall in
(other structures) or the first storey in the direction parallel to
0.0743
Ct = (wall structures) the
Ac
De hn ≤ 0.9

D
/ applied forces

Ac = Ae 0.2 + h e
n
� � � �2 ��

Empirical Method A Empirical Method B Ac(m2) = total effective area of the shear NZS 1170-5-(2004) (New Zealand)
T1 = 1.0kt h0.75
n serviceability limit state T1 = 1.0kt h0.75
n walls in the first storey of the building;
T1 = 1.25kt h0.75
n ultimate limit state n
serviceability limit state T1 = 1.25kt h0.75 Ai(m2) = effective cross-sectional area of ith
kt = 0.075 ultimate limit state shear wall in the first storey of the build-
0.0743 ing;
Ac
lwi(m) = length of ith shear wall in the first
kt = √

Ac = Ai 0.2 + lwi hn storey in the direction parallel to the


applied forces;
∑� � � � �2 ��

lwi hn ≤ 0.9
/

T1 = Ct h0.75
n n
1. T1 = Ct h0.75 Eurocode 8-Part-1 (2009): BS EN 1998-1
Ct = 0.075 (2004) (European Union)
Ct = 0.005 (other structures)
2. Same as Empirical Method B from
NZS 1170–5-2004
3∕4 3∕4 NBCC (2015) (Canada)
Ta = 0.075hn Ta = 0.05hn shear wall structures
Bulletin of Earthquake Engineering
Table 1  (continued)
Reinforced concrete frames Reinforced concrete and masonry wall Remarks Code/standard (country)
structures
1. Ta = 0.1NN ≤ 12 Ta = Ct hxn N = the number of storeys where the average ASCE 7-16: (2017) (USA)
2. Ta = Ct hxn story height is at least 10ft;
Where
Ct = 0.0466 AB(ft2) = area of base of structure;
x = 0.9 Ct = 0.0488 x = 0.75 Ai  ( ­ft2) = web area of ith shear wall;
(other structures), or Di(ft) = length of ith shear wall;
Ct = 0.0019 ,x = 1.0 hi(ft) = height of ith shear wall;
Cw x = number of shear walls in the building

Bulletin of Earthquake Engineering

(wall structures) effective in resisting lateral forces in the


Ai
hn
direction under consideration
Cw = 100
A h h
B i=1 i 1+0.83 Di
x � �2

i


� �2 �

Ta = 0.075h0.75 For all other buildings including moment d (m) = base dimension of the building at the IS1893 Part-I (2016) (India)
resisting frame buildings with masonry plinth level along the considered direction
infill wall panels of lateral force
Ta = 0.09h h = height of the building (in m)
d Aw = Total effective area ­(m2) of walls in the

For buildings with RC structural walls first storey of the building


0.075h0.75
Ta = ≤ 0.09h Nw = Number of walls in the considered
Aw d direction of lateral load
√ √

L
Aw = Awi 0.2 + hwi
i=1
Nw �
� � �2 ��

Awi = Effective area ­(m2) of wall-i in the

first storey
Lwi = Length (in m) of structural wall-i in
the first storey in the considered direc-
tion of lateral load
Lwi h ≤ 0.9
/

13
Bulletin of Earthquake Engineering

1.2 Objectives

Based on the limited review presented above and followed by the identified research gap,
this paper is aimed to address the following three key issues: (1) An empirical relation for
the natural period exclusively based on CM building stock; (2) Associated damping ratio;
and (3) Distribution of base shear along the height for force-based seismic design. AVT
will be used for the characterization of modal properties. Although the principle remains
same, the processing of AVT data for extracting modal characteristics varies in prior art
contingent on the size of the project and, number and type of the available sensors. Another
objective of this paper is to review some of the available frameworks of AVT and select the
appropriate one with due modifications, if needed.
Description of the CM building stock is first presented followed by that of the test setup.
Second, commonly used AVT frameworks are reviewed followed by the description of
adapted framework. Third, results of AVT are presented and analysed for the modal prop-
erties. Fourth, Welch method is used with some modification for estimating the damping of
CM buildings. Fifth, modal characteristics extracted using AVT are compared against the
numerical model, which is then tuned accordingly. Sixth, experimentally obtained natural
periods are used in developing empirical equation as possible recommendation for seis-
mic code. Finally, the tuned numerical models are analysed subjected to an ensemble of
recorded ground accelerograms to arrive at the distribution of base shear along the height.
This must be noted to avoid any possible misinterpretation that the present paper is nei-
ther aim to identify the possible location nor to assess the extent of damage in the (con-
fined) masonry buildings. Specialized literatures (for example, Valente and Milani 2018,
2019) may be referred to for that purpose. Extraction of modal characteristics can be per-
formed using a relatively simpler (auto- and/or cross-) correlation studies of the ambient
signatures. However, the framework adopted in this paper is relatively complex involving
the spectral value decomposition, merging strategies with rescaling, checking for modal
coherence etc. as the eigen shape of dominant modes are also one of the primary objec-
tives. Though free vibration analysis suffices for the numerical validation of the mode
shapes, the vertical distribution of base shear is investigated through linear dynamic analy-
sis of the tuned numerical model subjected to an ensemble of ground motion. Since seis-
mic code recommended vertical distribution of base shear in force-based design is intended
to capture the fundamental as well as higher mode effects in linear elastic structure rather
than that at the post-yielding stage with period elongation, more sophisticated analysis such
as nonlinear static pushover and nonlinear dynamic analysis are not included in this paper.
However, it would be interesting to explore the true behaviour of the confined masonry
buildings designed through the recommended natural period and vertical distribution of
base shear in a future work.

2 Building stock

The building stock includes a set of 6 hostel blocks at the residential campus of Indian
Institute of Technology Gandhinagar (Ahmedabad) located in Seismic Zone III (moder-
ate seismicity with peak ground acceleration (PGA) at maximum considered earthquake
(MCE) is 0.16 g). Bashir and Basu (2018) conducted probabilistic seismic hazard analysis
(PSHA) for the entire state of Gujarat. PGA (in g) associated with the 2475 years return
period (MCE) for Ahmedabad was reported as 0.10 (bedrock), 0.20 (hard soil) and 0.23

13
Bulletin of Earthquake Engineering

(medium soil) against that recommended by Indian Standard (IS 1893-Part-1 (2016)) as
0.16 regardless of the soil site. Note that IS 1893-Part-1 (2016) does not explicitly account
for the effect of soil sites through amplification/de-amplification of PGA but, recommends
adjustment of the right corner period instead. For the soil sites of hard to medium, PGA
associated with the MCE should have been in the order of 0.2 g. However, for the com-
pliance of Indian Standard, the building stocks are designed assuming a PGA at MCE of
0.16 g. Further, IS 1893-Part-1 (2016) recommends 50% of the PGA at MCE for the asso-
ciated design basis earthquake (DBE). In contrast, Bashir and Basu (2018) reported the
ratio of PGA in MCE to DBE for Ahmedabad as 1.76 (bedrock), 1.68 (hard soil) and 1.57
(medium soil). Once again, for the compliance of Indian Standard, the ratio of PGA in
MCE to DBE is chosen as 2 instead of that based on PSHA (which is about 1.5).
Each hostel building is G + 3 (four)-storied and is further divided into different wings.
These wings are isolated by expansion joints. Hence, each wing behaves as an isolated
structure and may have a lift and a staircase attached. The general layout consists of a rec-
tangular strip with a corridor through the middle and rooms on either side. However, some
wings may have irregular layouts other than the regular one. For the sake of uniformity,
the wings are chosen such that they remain by and large rectangular. Hence, nine of them
are finalized and AVT is performed on them. The layout of hostel buildings is presented in
Fig. 1 with the expansion joints marked by solid lines and the wings selected for AVT by
dotted enclosure. Each of these buildings is named for the convenience of data organiza-
tion. Typically, a building is denoted as X_Wk. Here, X denotes the name of the hostel,
for example, B, C, D, E or F whereas the Wk identifies the wing number, for example,
W1, W5 etc. Salient features of the geometry and structural aspects are summarized below:
(1) Wall thickness of 230  mm for rooms and 350  mm for common restrooms; (2) Slab
thickness of 150 mm; (3) Confining elements with a cross-section of 230 mm × 230 mm;
(4) Typical storey height of 3.2 m; (5) Concrete of M25 grade (characteristic compressive
strength is 25 MPa) having unit weight of 25 kN/m3 (IS 456 2000); (6) Fly ash masonry
unit (i.e., brick) with a compressive strength of 9.5 MPa and unit weight of 17 kN/m3; (7)
Design live load of 3 kN/m2 on floor as well as roof; and (8) Weight contributed from the
non-structural components is included through the floor finish with a distributed load of
1.25 kN/m2.
It should be clear that Nema and Basu (2019) considered three G + 2 staff and faculty
housing apartments located in the same campus. AVT was performed using only one sen-
sor and natural frequency and damping ratio were reported. Clearly, mode shapes could not
be constructed using one sensor. This paper adopts a completely different framework for
AVT of G + 3 hostel buildings using a set of six sensors as described in what follows.

3 Test setup

3.1 Sensors and data acquisition

The sensors used are Dytran-3191A1 with sensitivity of about 10 V/g and noise floor level
of 3 × 10–6 g. The maximum magnitude that it can measure is about 0.5 g which is suffi-
cient for AVT characterized by its low amplitude. These sensors are uniaxial with a trans-
verse sensitivity of 5% of axial sensitivity. Each sensor records vibration only in the axial
direction while the influence of transverse vibration is negligible. An acquisition module
and chassis from National Instruments (with 24-bit resolution and in-built anti-aliasing

13
Bulletin of Earthquake Engineering

Fig. 1  General hostel buildings layout with tested wings labelled

13
Bulletin of Earthquake Engineering

filter) are used for collecting the data. Data logging is performed using NI-LabVIEW soft-
ware, sampled at about 1.67 kHz and each sample record is taken for about 3 min. A total
of about 15 sample records are gathered for each setup (to be discussed later in this paper).

3.2 Sensor layout

All the four levels are mounted (using adhesive) with three sensors each. Six sensors are
available for this experimental program but the number of DOFs of interest are 12 (under
the assumption of in-plane rigidity of the floor diaphragm, 3DOFs per floor level) and
hence, the sensor referencing is used with three setups. Three out of six are used as the ref-
erence (stationary) sensors while the other three are used as the roving sensors. Reference
sensors are placed at the second floor while the others are placed at a different floor levels
in every setup. A typical sensor layout is presented for the floor level and that for roof level
(Fig. 2). At every floor level, two of the three sensors are deployed along the same direc-
tion (co-oriented) while the remaining is placed along the perpendicular direction. These
sensors together capture the in-plane DOFs under the assumption of in-plane rigidity of
floor diaphragm. All these sensors are attached to the wall at different locations but closer
to the respective floor levels.
The layout described above applies to all the buildings except E_W5. A different layout
is explored here to understand the bias, if any. Six sensors are placed in four setups with
three reference sensors remain stationary at the first floor. Two reference sensors are placed
to measure vibration along shorter dimension of the building (say x direction) and one sen-
sor along the longer direction (say y direction). Layout of the roving sensors in four differ-
ent setups are as follows: (1) First setup: only two sensors are placed along x direction in
second floor; (2) Second setup: only two sensors are placed along x direction in third floor;
(3) Third setup: two sensors are placed along y direction in third floor while one along the
same direction but in second floor; and (4) Fourth setup: two sensors are placed along x
direction while one along the y-direction in fourth floor.
It is always possible to adopt a better sensor layout with an aim to capture the torsional
modes. However, as explained later in Sect. 4.2.5, the torsional mode shapes will always
be in error if the mass distribution is unknown and nonuniform. Since primary objective
of this paper is to achieve the empirical equation for natural periods and the distribution of
design base shear along the height, two dominant translational modes are given emphasis
for the construction of mode shapes followed by the fine tuning of numerical model.

3.3 Stacking and minimum recording time

In order to reduce the noise in recorded data and hence, to achieve a relatively smoother
representation (of spectral densities or the associated singular values) in frequency domain,
stacking is often employed using a set of non-overlapping rectangular or overlapping
other (say, Hanning) windows. Rectangular stacking window is employed in this paper as
described below. Recorded array of time series is first divided into ten non-overlapping
time segments. Spectral density matrix is calculated for each segment and the associated
singular values are extracted at any sampled frequency. The respective singular values are
averaged over the number of stacking to arrive at a smoother representation over the entire
frequency range. However, while performing stacking ) operation, Brincker and Ventura
(2015) recommended a minimum recording time Tr,min in order to reduce the bias error in
(

estimating correlation functions:

13
Bulletin of Earthquake Engineering

Fig. 2  Sensor layout at typical floor and terrace, B_W1

13
Bulletin of Earthquake Engineering

N
Tr,min =
2𝜉fmin (1)

Here N is the number of non-overlapping stacking windows, 𝜉 is the damping ratio and fmin
is the minimum frequency of interest. For a minimum frequency of 4 Hz and 5% damping
ratio, the minimum recording duration for 10 stacking windows is about 25 s. Although the
data is recorded in this paper over 180 s, the number of stacking windows are restricted to
10 throughout the study to avoid the reduction in frequency resolution.

4 Ambient vibration test (AVT) framework

Salient details of some of the commonly used frameworks are reviewed here followed by
which is adapted in this paper.

4.1 Review of commonly used AVT framework

4.1.1 Time domain analysis

Two methods are popular under this category, namely, (1) Covariance-Driven Stochastic Sub-
space Identification (SSI-Cov) and (2) Data Driven Stochastic Subspace Identification (SSI-
Data) (Xie et al. 2016; Liu et al. 2013; Conte et al. 2008). The major difference between SSI-
Cov and SSI-Data is that covariance matrix is computed in the former while the latter involves
projecting row space of future outputs onto that of past outputs. Both are very robust and suit-
able for practical applications although SSI-Cov is faster than SSI-Data (Peeters and De Roeck
2001). These algorithms are parametric in nature and thereby relying on the underlying math-
ematical model (state space model).
Associated state-space representation (Heylen et al. 1995) of an N -DOF system takes the
form with order 2N  , which however may not apply to a stochastic state-space model (Peeters
and De Roeck 1999) owing to the finite data size, limited precision in computation, presence
of measurement noise and the idealization involved in the modelling. Since the model order is
unknown, stabilization is often recommended by comparing the modal estimates (frequency,
damping etc.) against the varying model order. Though the procedure can be used to iden-
tify the stable modes from the noise modes (characterized by spurious estimates), extracting
modal parameters for every model order may appear to be tedious.

4.1.2 Frequency domain analysis

Peak-picking is non-parametric in nature and perhaps the simplest approach under this cat-
egory for extracting the frequency and mode shape. Spectral density functions are plotted
against frequency and possible candidates of natural frequencies are identified though pick-
ing the peaks. Under the assumption of low damping and widely separated modes, the mode
shape can be obtained through the following relation
H
(2)
( )
Gyy 𝜔j ≈ 𝛼j uj uj

Here Gyy denotes the spectral density matrix (arrived at from the processing of recorded
response); j denotes the sequence of mode; 𝜔 and u denote the estimate of natural frequency

13
Bulletin of Earthquake Engineering

and mode shape, respectively;𝛼 is the associated scaling factor; and the superscript H
denotes the Hermitian. However, the estimated modal characteristic may lead to grossly
inaccurate inferences if the underlying assumptions are not strictly satisfied. In that case,
these mode shape estimates are referred to as operational deflected shapes instead, which is
a result of interaction between the closely spaced mode shapes. In order to overcome these
limitations, Frequency Domain Decomposition (FDD) was introduced (Brincker et  al.
2001; Bendat and Piersol 2011) making use of the singular value decomposition (SVD) on
Spectral density matrix.
The singular values are real-valued and usually arranged in the descending order while
the singular vectors have complex coefficients. One advantage is that the number of non-
zero singular values (subjected to an assumed condition number, ratio of the largest to
smallest singular values) determines the rank of spectral density matrix, which in turn indi-
cates the number of closely spaced modes contributing to the peak. Hence the singular val-
ues are also known as the Complex Mode Indication Function (CMIF). The mode shape is
given by the first singular vector (with rank of the matrix as unity) in case of widely sepa-
rated modes, which is similar to the peak picking method. If multiple modes contribute to
the peak, all singular vectors corresponding to non-zero singular values are considered as
the mode shapes. However, orthogonality bias is introduced because these singular vectors
satisfy geometric orthogonality whereas actual mode shapes are orthogonal with respect to
mass matrix (Brincker and Ventura 2015).
Another extension of FDD, namely, Extended Frequency Domain Decomposition
(EFFD) has an added functionality while estimating damping (Brincker et al. 2000). Over-
all, the FDD is known for being fast, simple, free from computational modes and can iden-
tify real modal estimates from harmonic ones (Zhang and Brincker 2005; Jacobsen 2006).
However, picking of mode shapes may be difficult in case of closely spaced or highly
damped modes. Since it involves some subjectivity, quality of mode shape depends on
the frequency picked within the resonance band and hence, the inference drawn may not
always represent the best estimate (Peeters and De Roeck 2001). Also, the estimated damp-
ing is not good when compared with other methods.

4.1.3 Referencing

In order to overcome the issues on availability of limited number of sensors, the reference-
based data acquisition was realized by Felber (1994). In referencing, total number of sen-
sors are divided into Reference Sensors and Roving Sensors. For every setup, the Reference
Sensors remain at the same location (also often referred to stationary) while the Roving
Sensors are shifted to different locations. Responses associated with all the required DOFs
can be extracted by adopting this strategy. Assumption of stationary white noise excita-
tion does not theoretically require the Reference Sensors. However, loading condition does
not remain stationary in practice owing to changes in environmental and operational con-
ditions, movement of pedestrian etc. Associated challenges can be overcome by identify-
ing the changes in loading spectra through Reference Sensors followed by rescaling of the
resulting mode shape coefficients.

4.1.4 Merging strategies

Post-Separate Estimation Rescaling (PoSER) is traditionally used to merge the mode shapes.
It involves estimation of mode shapes from each setup separately and merging them by

13
Bulletin of Earthquake Engineering

rescaling at different places, which makes the process tedious. Alternatively, the other merging
strategies like Post-Global Estimation Rescaling (PoGER) and Pre-Global Estimation Rescal-
ing (PreGER) (Parloo 2003; Parloo et al. 2003) may also be used for this purpose. Both the
merging strategies aim to analyze the data from multiple setups simultaneously and they are
implemented in time domain method such as SSI-Cov and SSI-Data, and also in some vari-
ants of frequency domain method such as Maximum Likelihood Estimator (Parloo et al. 2003;
Reynders et  al. 2009). However, implementation of such strategies has not been explicitly
reported for FDD method although Peeters and De Roeck (2001) realized that FDD could also
be used along with these merging strategies.

4.2 Adapted AVT framework for natural frequencies and mode shapes

4.2.1 Preprocessing

For a given building, data from all the setups are taken together for the purpose of analysis.
Before estimating spectral density matrix, data is checked for any anomalies like clipping or
response due to large impact loading by checking their autospectral density functions.

4.2.2 Response at degrees of freedom extracted from recorded data

Under the assumption of in-plane rigidity of floor diaphragm, rotational (torsional) DOF at
any floor level is extracted from the translational responses simultaneously recorded using
two uniaxial sensors oriented along the same direction (co-oriented). In other words, the time
series associated with the floor rotational DOF can be computed as
u1 − u2
𝜃= (3)
L
where u1 and u2 are the simultaneously recorded time series from a pair of uniaxial sensors
deployed at different locations of the same floor along the same direction and L is the dis-
tance between the locations perpendicular to the direction of measurement. For simplicity,
the direction of measurement is taken along the direction of one of the translational DOFs.
On the other hand, if a pair of triaxial sensors are used, it is recommended to rotate the
recorded horizontal pair of time series at each location in order to extract the component
along the direction normal to the line connecting the stations. These two time series can be
used in Eq. (3) with L as the separation distance while extracting the rotational time series.
This way of measuring rotation is common in rotational seismology (Basu et al. 2017) but
not used in the present paper as only uniaxial sensors are available.
Although not mandatory, it is often preferred to consider the DOF at a point close to the
geometric center of gravity (C.G) at respective floor levels. Such a location however may not
always be accessible for the deployment of sensors. In such a case, sensors are deployed at
other locations and recording at the reference point (geometric c.g.) in respective floor level
may be extracted as
uxref = ux + ry 𝜃
(4)
uyref = uy − rx 𝜃

13
Bulletin of Earthquake Engineering

Here ux uy denotes the recording at station ry rx distance away from the reference point
( ) ( )

along the direction y(x) and θ is the floor rotation (torsional) measured in counter-clock-
wise direction.
In this paper, the number of uniaxial sensors deployed at any floor level is three, namely,
two along the same translational DOF while the third one along other translational DOF.
Therefore, the present instrumentation extracts the signal at the reference point along the
DOFs through an ‘exactly-determined’ problem. On the other hand, if triaxial sensors are
used or even more than three uniaxial sensors are used at any floor level, one will have to
treat an over-determined problem instead. This investigation is beyond the scope of present
paper and will be reported separately elsewhere.

4.2.3 Merging strategy

Different merging strategies are generally used while combining the recorded data of refer-
ence and roving sensors from different trials of experiment. PoGER and PreGER are the
two most used strategies as briefly explained below.

4.2.3.1  PoGER 

1. For every ith setup, the spectral density matrix (or the correlation matrix) given by
Gy,yref (𝜔)| (or Ry,yref (𝜏)|  ) is computed. The cross spectral density terms are computed
| |
|i |i
between the total( number ) of ( sensors and the reference sensors and hence, the matrices
are of the size ns + nri × ns  . Here ns and nri are the number of reference and roving
)

sensors, respectively, used in the setup.


2. A global spectral density matrix Gyy (𝜔) is next formed by stacking individual matrices
g

on top of each other for k number of setups as shown below:

⎡ Gy,yref (𝜔)�1 ⎤
⎢G (𝜔)�2 ⎥⎥
Ggyy (𝜔) = ⎢ y,yref (5)
⎢ ⋮ ⎥
⎣ y,yref (𝜔)�k ⎦
⎢G ⎥

3. This spectral density matrix is then analyzed, and mode shape vector is rescaled to
ascertain the final mode shape. Rescaling of mode shapes is explained later in this paper.

4.2.3.2  PreGER  PreGER strategy also has the same first step but rescaling is performed
on the spectral density matrix rather than the mode shapes. One may refer to Parloo et al.
(2003) for further details of the procedure.
Spectral density matrix is rectangular in both the strategies and hence, the interpretation
of its singular value decomposition (SVD) is important. For example, the matrix Gyy (𝜔)
g

may be decomposed as follows.


[ ]
Ggyy (𝜔) = [U]Ntot ×Ntot [S]Ntot ×ns [V]Tn ×n (6)
N ×n s s
tot s

Unlike FDD, the right singular vectors deliver a condensed form of mode shape for an
over-determined problem that does not have a physical meaning. On the other hand, the left
singular vectors correspond to an underdetermined problem.

13
Bulletin of Earthquake Engineering

Selection of merging strategy is somewhat subjective and often contingent of the experi-
mentalist’s preference, available resources, accessibility, and sensor layout. Final results may
not differ appreciably unless a gross mistake is committed during the data acquisition and pro-
cessing. A slightly different merging strategy is adopted in this paper that caters better physi-
cal insights as explained below.

4.2.3.3  Adapted merging strategy 

1. All the reference and roving sensors data different setups are assumed to be obtained
from a single setup. Responses at the DOFs are then extracted and taken together while
computing the spectral density matrix Gyy . This is of the size Ntot × Ntot and used in
FDD.
2. Mode shape vectors are next extracted through SVD as explained in the subsequent
section, which is then rescaled to obtain the final estimate.

It is similar to PoGER but differs in the issues that (1) Cross Spectral Density (CSD) of
roving sensors within a setup is taken into account; and (2) CSD also considers the DOFs
across the setups. Nonstationarity of excitation is accounted for through the response recorded
at the reference sensors followed by appropriate rescaling of the mode shapes. A comparison
of different merging strategies is beyond the scope of this paper and will be addressed else-
where whenever available.

4.2.4 Estimating modal parameters from the response due to non‑stationary ambient


excitation

Often the excitation is tacitly assumed as stationary during operational modal analysis which
however has not been the case in this paper. This section provides a theoretical basis for such
relaxation as follows. Denoting y(t) as the time series of response vector recorded at all the
DOFs due to a nonstationary ambient excitation field and q(t) as the associated vector of
modal coordinates, the time and frequency domain relation can be expressed though modal
matrix Φ as

y(t) = Φq(t) and Y(f ) = ΦQ(f ) (7)

Here Y(f ) and Q(f ) denote the Discrete Fourier Transform (DFT) of y(t) and q(t) , respec-
tively. Note that unlike Fourier Transform,y(t) and q(t) are not considered to be stationary
for the evaluation of DFT and instead, it is sufficient to assume the slowly varying fre-
quency modulation at the respective frequency bands. The recorded time series associated
with jth DOF can be written as
∑ ∑
yj (t) = yjk (t) = mjk (t)y0k (t) (8)
k k

Here k denotes the frequency band, m is the associated slowly varying modulating function
with y0k as the respective stationary component. Under the assumption of slowly varying
modulation, taking Fourier Transform on either side, Eq. (8) leads to
∑ ∑
Yj (f , t) = Yjk (f , t) = mjk (t)Yk0 (f ) (9)
k k

13
Bulletin of Earthquake Engineering

Nonstationarity in the frequency content at this stage may be noted. However, the ambi-
ent loading condition may not abruptly change during the period of data acquisition. In
such a case, the modulating function in each frequency band may be considered to remain
constant over the recording duration. Consequently,
∑ ∑
mjk Yk0 (f ) = DFT yj (t)
[ ]
Yj (f ) = Yjk (f ) ≈ (10)
k k

Note that the assumption of slowly varying frequency modulation does not offer any
limitation. Alternatively, one may use Hilbert Transform without such assumption as
explained in Rodda and Basu (2020) but in a different context. In such a case, one will have
to deal with an evolutionary spectral density matrix. Nevertheless, the extracted natural
properties will turn out to be time invariant. Hilbert Transform however is not used in this
paper.
Now, returning to the modal expansion given by Eq. (7) and denoting the superscript H
as the Hermitian, one may write
1 H 1 H 1 H
[ ] [ ] [ ]
E Y(f )Y (f ) = E ΦQ(f )Q (f )ΦH = ΦE Q(f )Q (f ) ΦH
T T T (11)
Gyy (f ) = ΦGqq (f )ΦH

here Gyy and Gqq denote the spectral density matrices of the recorded response and the
associated modal coordinates; and E[.] denotes the statistical expectation over an ensemble.
Note that for real mode shapes, Hermitian operation is equivalent to taking transpose. Tak-
ing SVD of the spectral density matrix computed in Sect. 4.2.3 (under the adapted merging
strategy),

Gyy (f ) = USU H (12)

Here S is a diagonal matrix comprising of the singular values while U is spanned by the
singular vectors. Apparently, considering the equivalence of Eqs. (11) and (12), the singu-
lar values may be viewed as the spectral density functions of modal coordinates while the
singular vectors may be related to the mode shapes. Singular values are generally sorted in
descending order after SVD. If the first singular value is plotted against frequency and the
peaks are observed, the first singular vector corresponding to a given peak may be related
to the associated mode shape, as explained below.

4.2.4.1  Relating singular vector to mode shape  By the assumption of uncorrelated modal
coordinates, i.e., Gqq is diagonal, modal decomposition can be performed by spectral value
decomposition and hence, the equivalence of Eqs. (11) and (12) can be expressed as

U(f ) S(f ) dia U H (f ) = Φ Gqq (f ) dia ΦH


[ ] [ ]
(13)

Taking expansion on either side,


N N
H
∑ ∑
sii (f )ui (f )ui (f ) = gqq (f )Φq ΦH
q (14)
i=1 q=1

Given a peak at f =(f0 ,)that can (correspond to any one of the modal coordinates, say jth,
and in such a case, gjj f0 >> gqq f0 for q ≠ j . Since singular values are (usually (based
)

on the algorithm used, say, MATLAB) arranged in descending order, s11 f0 >> sii f0
) ( )

13
Bulletin of Earthquake Engineering

for i ≠ 1 . Therefore, considering the contribution from dominant term only (and thereby
neglecting the remaining terms) on either side, one may write
( ) ( ) H( )
s11 f0 u1 f0 u1 f0 = gjj f0 𝜙j 𝜙H
( )
j
( ) (15)
u1 f0 ∝ 𝜙j

Therefore, singular vectors in general do not represent the mode shapes. For example,
at any sample frequency that may or may not represent one of the natural frequencies,
there exists a set of singular vectors satisfying the orthogonality condition U(f )H U(f ) = I  .
In contrast, at any natural frequency, there exists only one mode shape and the modal
matrix when constructed through all the mode shapes satisfies the orthogonality condition
ΦT MΦ = I  . However, only at the peak of first singular value, mode shape can be estimated
as the respective first singular vectors. Hence, it is necessary to identify possible candidates
of natural frequency for correctly identifying the mode shape as discussed in Sect. 4.2.6.

4.2.5 Post‑processing

4.2.5.1  Rescaling of mode shapes  After extracting an estimate of mode shape (i.e., the first
singular vector) 𝜙 of size Ntot , rescaling is performed as follows:
Consider a total of k datasets with a given number of reference sensors. The mode shape
estimate 𝜙 can be written as
]T
𝜙 = 𝜙�1 𝜙��1 𝜙�2 𝜙��2 ⋯ 𝜙�k 𝜙��k (16)
[

here the mode shape coefficients contributed from the ith dataset are divided into two cate-
gories: 𝜙′i and 𝜙′′i denote the DOFs associated to reference and roving sensors, respectively.
Compare all the 𝜙′i , i = 1, k and pick the one comprising of the largest (regardless of sign)
mode coefficient. Let jth setup contains the largest mode shape coefficient in its lth entry
and denote it as 𝜙′j,l . Consider only the jth setup for the DOFs associated with reference
sensors and the resulting mode shape can be expressed as
[ ��
]T
𝜙 = 𝜙�j 𝜙��2 𝜙2 ⋯ 𝜙��j ⋯ 𝜙��k (17)

The mode shape can be further normalized with respect to 𝜙′j,l or any other coefficient
deemed appropriate.

4.2.5.2  Modal coherence  The estimated mode shape that does not vary much around the
resonating frequency represents indeed a physical mode. If the peak is due to noise, the
resulting mode shape is expected to be irregular with high degree of fluctuation against a
minor perturbation in frequency. Modal coherence as defined below is estimated to assess
the stability of resulting mode shape before final acceptance and the implementation process
is adapted from Brincker et al. (2007) with some modifications:
| H ( )|
|u1 (f )u1 fi |
di (f ) = (18)
| |
| H || H ( ) ( )|
|u1 (f )u1 (f )||u1 fi u1 fi |
| || |
here u1 (f ) and u1 fi are the first column of the singular vector in U at any given frequency
( )

f and the ith peak frequency, respectively. If the following two conditions are strictly

13
Bulletin of Earthquake Engineering

satisfied, namely, (1) mass distribution at all the floor levels is identical (which may not
differ significantly in most buildings) and (2) instead of torsional rotation, its product with
radius of gyration at the respective floor levels are considered as the representation of the
rotational DOFs, the resulting mode shapes can be (proved as geometrically orthogonal
(without the weightage of lumped mass). Hence, di fj = 0 when fj is the jth peak fre-
)

quency and j ≠ i . However, owing to the (1) presence of instrument noise, (2) difference in
assumed and actual mass distribution, and (3) choice of torsional rotation (rather
( ) than mul-
tiplying with respective radius of gyration) as the DOFs, modal coherence di fj may not be
calculated as zero but a trough can be observed indicating a minimum.
Therefore, the modal coherence can be plotted
( ) against frequency for all the peak fre-
quencies followed by a check to ensure that di fj is maximum for i = j and minimum i ≠ j .
This is expected to identify the translational mode with reasonable accuracy, but the result-
ing torsional modes will be erroneous owing to the reason explained above. Nevertheless,
this is expected to affect neither the aim nor the inference of the present paper as explained
later.

4.2.5.3  Processing of complex mode shapes  Mode shapes identified from the framework
described above are complex (rather than real) implying a case of non-proportional damp-
ing. When an equivalent proportional damping is expected, as generally be the case with
seismic design, the complex modes must be converted into the associated real modes. The
technique used in this paper is adapted from Rainieri and Fabbrocino (2014) and explained
below. First, the complex modes are normalized with respect to the complex mode shape
coefficient of maximum magnitude. / For example, if Ψ is a mode shape given by the coeffi-
n
cients 𝜓1 𝜓2 … 𝜓n  , then Ψ → Ψ max 𝜓i . Next, the resulting complex coefficients are
[ ]
i=1
rotated towards the positive real axis if the real part is positive or towards the negative real
axis otherwise. In other words, the mode shape a + ib is converted into sgn(a)|a + ib|.

4.2.6 Tool for post‑processing

Picking of mode shapes is performed after the post processing. This is convenient since
all the data can be processed at once and visualized later. Post-processing is carried out by
a GUI program developed using Matlab, which displays the singular value and the mode
shape at any selected frequency. Modal coherence is also computed while picking in real-
time and are plotted along. The mode shape is finalized if there is high modal coherence
beyond certain threshold at and around the picked frequency. The GUI setup is shown in
Fig. 3 where mode shape coefficients, modal coherence, and singular value in both normal
and semi-log scale are plotted at a given frequency.

4.3 Adapted AVT framework for damping ratio

Damping ratios are estimated from the autospectral density functions (ASD) of accel-
eration time series recorded by each sensor at different locations. The procedure is
briefly described below: (1) Band-pass filtering of the time series around the resonating
frequency; (2) Estimation of ASD using Welch method and followed by inverse Fou-
rier Transform for computing the biased correlation; (3) Division of biased correlation
through a triangular window for the unbiased (correlation) Welch estimate (Brincker
and Ventura 2015); (4) Estimation of damping ratio assuming this correlation as the rep-
resentative of logarithmic decay in free vibration. Further details on the implementation

13
Bulletin of Earthquake Engineering

Fig. 3  GUI developed in MATLAB for post processing

Fig. 4  Algorithm used for the estimation of damping. Adapted from Nema and Basu (2019)

13
Bulletin of Earthquake Engineering

Fig. 5  Sample illsustration on unbiased correlation and the associated logarithmic decay

is presented elsewhere (Nema and Basu 2019). However, the necessary algorithm is
reproduced in Fig.  4 for the completeness and ready reference. Figure  5 illustrates a
sample case of unbiased correlation with units in some representative scales. It is rec-
ommended to include all the successive peaks as long as the clear trend of logarithmic
decay is available while estimating the damping ratio. For example, 5 peaks should be
considered with reference to Fig. 5.
All the sensors may not necessarily capture all the frequencies contingent on the sen-
sor orientation and input excitation. Hence, application of band-pass filtering around a
specified resonating frequency may also lead to unrealistic time series in some of the
sensors. Such cases are not considered here for the estimation of damping ratio. Further,
damping estimated from ambient signature represents only the inherent structural damp-
ing which is in fact a fraction of the actual damping available in the seismic signature of
a building. A major portion of the total damping is contributed from the system hyster-
esis on account of inelastic excursion. Therefore, damping estimated from the ambient
signatures does reflect the true scenario during a seismic event.

5 Illustrations on a chosen building

Figure  1 identifies the buildings considered in this experimental program. Only one
building, namely, B_W1 is considered here for the purpose of illustration, and plan lay-
out of its typical floor and that at the terrace are presented in Fig. 2.

Table 2  Distances of sensors Location X1 (mm) X2 (mm) Y1 (mm)


from reference locations in
building B_W1
Floor 1 130 3030 1780
Floor 2 300 2390 2100
Floor 3 330 1130 1910
Terrace 0 1000 1100

13
Bulletin of Earthquake Engineering

5.1 Sensor layout

Sensor layout at a typical floor level and that at the terrace are given in Fig. 2 for the build-
ing B_W1. Sensor locations are marked by a cross ‘X’ with distance measured from the
respective reference points (Table 2).

5.2 Estimating frequencies and mode shapes

The building B_W1 is analyzed and the resulting singular values are calculated. Num-
ber of DOFs (under the assumption of in-plane rigidity of floor diaphragms) is 12
against the number of measured locations of 18 (six sensors deployed in three setups
with three sensors kept stationary for referencing). Spectral density matrix computed
based on all the 18 sensors are used for SVD and the resulting singular values are plot-
ted against frequency (Fig. 6). Similar data from the individual setups (with six sensors)
are presented in Fig.  7. It is instructive to compare both the Figs.  6 and 7: resolution
of the first (or first few) singular values are similar in all the cases. Therefore, increas-
ing the number of setups (or even the sensors) does not improve the identification of
natural frequencies through peak picking. This however has significant influence on the
resulting mode shapes. Nevertheless, only the first eight singular values are plotted in
Fig. 8 and the first eight frequencies are listed in Table 3. Modes 1–3 correspond to the
fundamental modes (first triplet) representing the lateral-torsional coupling (4–5  Hz),
as explained in the next paragraph. Modes 4–6 (7–10 Hz) might be due to the influence
of secondary structures, such as staircase, weakly connected to the primary structure
which is confirmed later in Sect.  5.4 through numerical model. Secondary structural
influence is also documented in the prior art (Goulet et  al. 2013; Au and Feng-Liang
2012). Low-rise confined masonry buildings are expected to deform in shear mode

Fig. 6  Singular values plotted against frequencies for the building B_W1

13
Bulletin of Earthquake Engineering

Fig. 7  Singular values of each setup plotted against frequency for the building B_W1

Fig. 8  First eight singular values plotted against frequency for the building B_W1

13
Bulletin of Earthquake Engineering

Table 3  Frequency estimates for Mode 1 2 3 4 5 6 7 8


the building B_W1 using AVT
Freq (Hz) 4.01 5.05 5.35 7.02 8.17 10.08 11.43 12.61

Fig. 9  Comparison of first three mode shapes of building B W1—AVT with numerical model (SAP 2000)

(Basu 2019) and natural frequencies in a purely shear beam are in the ratio of 1:3:5 etc.
(Humar 2012). Therefore, the second triplet of torsionally coupled natural frequencies
are expected to be in the range of 12–15 Hz, which is nearly confirmed by the modes 7
and 8 (11–13 Hz).
Mode shapes are computed using the procedure described above in Sect.  4. First
three modes are presented in Fig. 9 and the associated modal coherences are shown in
Fig.  10. Modal coherence for a physical mode is expected to be close to unity at and
around the picked natural frequency and vanishing at other natural frequencies. Each of
these three modes (after converting the complex coefficients into real) is first normal-
ized with respect to the respective maximum coefficient and a typical mode shape is
plotted separately along two horizontal and for torsional rotation. Consider, for exam-
ple, the mode shape coefficients along x-direction. These coefficients are not expected
to change the sign floor wise in any of the first three modes. Similar expectation holds
for the coefficients along y-direction and torsional rotations, as well. If this theoretical
expectation is degenerated to the dominant lateral directions/DOFs of vibration instead
of all three DOFs, Fig.  9 confirms the expectation in two out of three modes, namely
the first and third. Maximum coefficient is along y-direction in the first mode whereas
along x-direction in the third mode. The second mode is dominated by torsional vibra-
tion and hence, is expected to be in error owing to the reason discussed in context with
the modal coherence in Sect.  4.2.5. Further, this error also corrupts the mode shape
coefficients along the lateral directions also in the same mode.

13
Bulletin of Earthquake Engineering

Fig. 10  Modal Coherence plotted


against frequency for the first
three modes of building B_W1

13
Bulletin of Earthquake Engineering

It must also be noted that Fig.  8 indicates the presence of other modes, for example,
around 15 Hz and 25 Hz, but in the form of spikes which are not present in the numerical
model (presented later in Sect. 5.4). The 15 Hz and 25 Hz modes are probably contributed
from the exhaust fan running typically at 900 rpm and cooler’s fan running at 1300 rpm,
respectively. This is also confirmed with the modal coherence that indicates the resulting
mode shapes are randomly fluctuating within a narrow frequency band (0.25 Hz).

5.3 Estimating damping ratios

Damping ratio is estimated only at the triplet of torsionally coupled fundamental modes
using the procedure described in Sect. 4.3. Recorded time series may capture one, two or
all three frequencies of the triplet of fundamental mode contingent on the orientation of the
sensor deployment with respect to the orthogonal axes (if exists) of the building and energy

Fig. 11  Auto spectral density before filtering at different locations in the building B_W1

13
Bulletin of Earthquake Engineering

Fig. 12  Auto spectral density after filtering (3.6–6 Hz) at different locations in the building B_W1

distribution of the ambient loading to excite the entire triplet. Signal recorded by each sen-
sor is considered and the ASD is constructed employing the stacking of 10 nonoverlapping
rectangular windows. Figure 11 presents the results for all the 18 sensors. A narrow band-
pass filter (3.6–6  Hz that encompasses the triplet of fundamental mode) is employed on
the recorded signals and the resulting ASDs are shown in Fig. 12. Since one sensor may
capture one or more frequencies of the triplet, the frequency (ies) captured by each sensor
is (are) noted. Recording of a typical sensor is then considered, a narrow band-pass filter
centered on one of the captured frequencies is implemented and the resulting time series
is extracted. ASD is then estimated using Welch method. Each sensor has captured only
one frequency in this building, which however, may not always be the case in practice. In
such a case, one sensor may lead to more than one ASD using Welch method. The result-
ing correlation functions (as discussed in Sect.  4.3) normalized with respect to its peak,
are shown in Fig. 13. Estimated damping ratios are presented in Table 4 (second column),
which vary form 2–9% with a mean of 3.6% and COV as 0.6. Note that each sensor has led

13
Bulletin of Earthquake Engineering

Fig. 13  Normalized autocorrelation function marked with peaks for sensors at different locations in B_W1

to one damping ratio for this building. Finally, it must also be noted that the damping ratio
computed from AVT may not represent the true scenario during a seismic excitation as
discussed in Sect. 4.3.

5.4 Numerical model and fine tuning

5.4.1 Macro modelling

Basic modelling of the structural elements like frame, wall, slab, and staircase is discussed
below followed by the weak link and meshing. Material definitions used for these elements
are provided in Table 5. Young’s modulus of concrete is considered per Indian Standard (IS
456 2000) with Poisson’s ratio as 0.3. Compressive strength and Young’s modulus adopted
for masonry have the basis of tests performed onsite during construction and in laboratory.

13
Bulletin of Earthquake Engineering

Table 4  Damping ratio (% of critical) estimated using AVT for the entire building stock
Set B_W1 B_W2 B_W3 C_W1 C_W2 C_W4 E_W4 E_W5 F_W2

1 4 5 7 2 6 6 6 4 4 5 4 4
2 2 3 6 3 4 6 3 6 5 5 5 4
3 2 4 6 4 5 7 6 4 4 5 5 4
4 5 5 7 2 10 7 6 4 4 5 4 4
5 2 3 6 3 4 6 4 4 4 5 5 3
6 2 5 6 3 7 7 6 4 4 7 5 4
7 9 4 6 3 6 6 6 4 4 5 4 5
8 2 4 6 3 3 6 3 4 5 6 5 4
9 2 4 7 3 4 6 6 4 4 7 5 4
10 8 4 5 3 7 4 6 4 4 7 4 4
11 2 4 6 3 3 5 4 4 4 5 5 4
12 2 4 6 3 6 5 6 4 4 8 5 4
13 5 4 7 3 6 5 5 4 4 5 5 4
14 2 4 6 3 3 6 3 4 5 8 5 3
15 2 4 6 3 5 6 6 4 4 9 4 4
16 6 4 6 3 9 5 5 4 4 9 5 4
17 2 3 8 3 3 6 4 4 5 5 5 3
18 6 3 7 3 8 4 5 4 4 5 5 3
19 – – – – – – – – – 5 – –
20 – – – – – – – – – 6 – –
21 – – – – – – – – – 6 – –
22 – – – – – – – – – 5 – –
Mean 3.6 3.9 6.3 2.9 5.5 5.7 5.0 4.1 4.2 6.0 4.7 3.8
COV 0.6 0.2 0.1 0.1 0.4 0.2 0.2 0.1 0.1 0.2 0.1 0.1

Table 5  Material properties used in the numerical model of the building stock


Material Properties M25 Concrete Fly-Ash Masonry Fe500 Rebar

Compressive Strength (MPa) 25 3.2 500


Young’s Modulus (Gpa) 25 2.09 200
Unit weight (kN/m3) 25 17 –

Acceptance criteria for the batching of bricks is based on a minimum compressive strength
of ~ 9.5  MPa. Verma and Basu (2017) reported the laboratory test results correlating the
compressive strength of brick and prism strength of masonry. Latter was noted to be about
one-third of the former and hence, compressive strength of masonry is considered here as
3.2 MPa. Though Indian Standard (IS 1893-Part-1, 2016) recommends linearity of modu-
lus of elasticity of masonry with its prism strength, Verma and Basu (2017) reported the
linearity with square of the prism strength instead. Therefore, in line with the recommen-
dation of the Indian Standard, this paper assumes a linear relationship of modulus of elas-
ticity with the prism strength (not square!) but with a slightly higher slope so as to comply

13
Bulletin of Earthquake Engineering

with the test results: Young’s modulus of masonry is chosen as approximately 650 times
(against 550 times per Indian Standards!) of its compressive strength. Since the AVTs are
carried out under operational/functional state, the mass of the non-structural components
is included in the numerical modeling through that imposed over the floor area in terms of
floor finish.

5.4.1.1  Frame elements  Frame elements are modelled using line elements (three trans-
lational and three rotational degrees of freedom at each node) with Reinforced Concrete
section of M25 grade concrete (characteristic compressive strength as 25 MPa) and Fe500
grade steel bars (yield strength of 500 MPa) though the elastic analysis is based on gross
cross-section. These elements include the RC tie (columns and beams) of 230 mm × 230 mm
cross section and other beams and columns, if any.

5.4.1.2  Wall and slab elements  Both wall and slab components are modelled using 4-noded
shell elements with three translational and three rotational degrees of freedom at each node.
Slab is considered of 150 mm thick with M25 grade of concrete. Two types of wall elements
are used, namely, Fly-Ash Masonry of 230 mm and 350 mm thick. Since slab is modelled
explicitly, the floor diaphragm is not constrained with in-plane rigidity.

5.4.1.3  Staircase  Two types of staircase are modelled. One has a rectangular layout sup-
ported by shear walls on the sides and connects all the four storeys. The other one has a
circular layout with semi-circular landing supported by shear wall at the centre and connects
three storeys. The columns supporting the staircase are also heavier than the vertical ties
used elsewhere. The connection of staircase units with the main confined masonry building
needs special consideration as discussed below.

5.4.1.4  Weak link  The purpose of weak link is to connect the primary structure (building
wing) with the secondary structure (staircase or shear wall of lift). By architectural layout,
the secondary structure imparts a wide duct all throughout the height of building while the
primary structure is connected along its periphery. Gravity load shared by the secondary
structure is negligible when compared to that of the primary structure and seismic force
associated with the mass of primary structure is not assumed to be transferred to the sec-
ondary structure. Therefore, secondary structure is not included in the lateral load resisting
system and its connection to the primary structure is often considered as a weak link (or
even utilized as expansion joint, which however, is not the case here).
These secondary structures may induce additional frequencies to the overall structure
and influence the fundamental frequency of the primary structure. Weak link is modeled
here by reducing the stiffness of the connecting beam by an order of 1­ 0–3 of the original
stiffness.

5.4.1.5  Meshing  Rectangular mesh with size not exceeding 80  cm is used for shell ele-
ments when modelling the wall and floor slabs. Frame elements are also discretized accord-
ingly to accommodate the additional nodes on account of meshing the walls and slabs.

5.4.2 Comparing models with and without staircase

Figures  14 and 15 show the building (numerical model) without and with the stair-
case, respectively. Modal analysis is performed for both the models and the resulting

13
Bulletin of Earthquake Engineering

Fig. 14  Numerical model (SAP 2000) of the building B_W1 without staircase

frequencies are compared (Table 6). Note that floor diaphragm is not constraint as rigid
in its own plane and the slab is modeled explicitly as shell elements instead. Owing
to this, modeling of staircase connected to the primary structures through weak links
induces additional frequencies apart from slightly affecting the fundamental frequen-
cies, as evident from Table  6. Therefore, final choice of the stiffness of weak link is
likely to change the triplet of fundamental frequency too some extent.

5.4.3 Fine tuning of numerical model and comparison with experimental results

Numerical model that includes staircases (Fig. 15) is fine-tuned by adjusting the stiff-
ness of weak links and followed by a modal analysis. Fine tuning should also aim to
address the spatial variability of the material properties which however has not been
adopted here for simplicity. First triplet of natural frequencies of numerical model are
compared with the experimental results in Table  7 and close resemblance is noted.
Comparison of the associated mode shapes is presented in Fig.  9. While mode shape
coefficients along the dominant lateral direction of vibration match closely in experi-
mental observation and numerical model, such is not the case with the mode predomi-
nantly vibrating under torsion. This is expected as discussed in Sect.  4.2.5. Mode
shape coefficients along the lateral direction normal to the dominant direction of vibra-
tion also do not show well resemblance owing to the error in capturing the mode shape
coefficients associated with torsional rotation. This however is not a serious concern
for the main objective here. It must be noted that the first triplet of mode shapes does
not change noticeably and rather remain same after the fine tuning.

13
Bulletin of Earthquake Engineering

Fig. 15  Numerical model (SAP 2000) of the building B_W1 with staircase

Table 6  Frequencies from the Model Frequency (Hz)


numerical model of building
B_W1—without staircase (i.e.,
Without Staircase 4.28 4.70 5.30 7.38 – – 13.07
flexibly connected) and with
(flexibly con-
staircase (i.e., rigidly connected)
nected)
With Staircase
(rigidly connected) 4.40 4.75 5.24 7.21 7.83 10.11 11.90
% Difference 2.7 1.1 − 1.1 − 2.4 – – − 9.8

13
Bulletin of Earthquake Engineering

Table 7  Comparison of the first three natural frequencies of the building B_W1—AVT (Exp) and numeri-
cal model (SAP 2000)
Mode-1 Mode-2 Mode-3
Exp SAP2000 % Error Exp SAP2000 % Error Exp SAP2000 % Error

Freq (Hz) 4.10 4.405 7.4 5.05 4.75 − 5.9 5.35 5.24 − 2.1

6 Summary of results from other buildings

While the building B_W1 is chosen in the previous section for the illustration of the frame-
work adapted, eight other buildings, namely, B_W2, B_W3, C_W1, C_W2, C_W4, E_W4,
E_W5 and F_W2 are also considered in the AVT program. Numerical modeling of these
buildings is also carried out using the same principle as discussed earlier. Details of AVT
procedure, for example, plan layout, instrumentation etc. are omitted here for brevity. Simi-
larly, description of the numerical model is also not discussed here. These details have
been presented elsewhere (Chakra-Varthy 2019). This section only summarizes the main
results of AVT and numerical modeling.
Table 8 presents the comparison of the first triplet of fundamental frequencies extracted
from AVT of all the nine buildings and respective numerical models. Maximum error for
each building is also highlighted. Out of 27 natural frequencies, the maximum error is
noted as 16% in one case and in all other cases, it is less than 8%. This resemblance may be
considered acceptable for all practical purposes. Similarly, comparison of mode shapes is
presented in Fig. 16 only along the dominant translational directions for the all the build-
ings other than B_W1 (Fig. 9). Once again, the resemblance in experimental observation
and numerical model is noted as satisfactory.
Damping is estimated only for the first triplet of torsionally coupled modes using the
procedure described in Sect. 4.3 and the results are summarized in Table 4. In some of the
buildings, for example, B_W2, C_W2 and F_W2, each sensor has captured two of the first
triplet of fundamental frequencies and hence, the results indicate two columns of damping
ratio. Distribution of damping ratios is shown in the histogram of Fig. 17. Median damp-
ing is noted as 5%. This is also close to what reported by Nema and Basu (2019) based on

Table 8  Comparison of the first three natural frequencies of the entire building stock—AVT (Exp) and
numerical model (SAP 2000)
Building Exp SAP2000 % Error Exp SAP2000 % Error Exp SAP2000 % Error

B_W1 4.1 4.4 − 7.3 5.05 4.75 5.9 5.35 5.24 2.1
B_W2 4.07 4.3 − 5.7 5.07 4.85 4.3 5.44 5.39 0.9
B_W3 4.04 4.26 − 5.4 4.77 4.61 3.4 5.21 4.94 5.2
C_W1 4.14 4.25 − 2.7 4.51 4.57 − 1.3 5.08 4.74 6.7
C_W2 4.09 3.95 3.4 4.54 4.4 3.1 5.07 5.06 0.2
C_W4 3.94 4.58 − 16.2 4.96 5.37 − 8.3 6.38 6.09 4.5
E_W4 3.99 4.29 − 7.5 4.89 4.63 5.3 5.44 5.39 0.9
E_W5 4.22 4.29 − 1.7 4.61 4.57 0.9 5.14 5.24 − 1.9
F_W2 4.42 4.65 − 5.2 5.09 4.92 3.3 5.55 5.46 1.6

13
Bulletin of Earthquake Engineering

Fig. 16  Translational mode shape along the dominant directions in different buildings

13
Bulletin of Earthquake Engineering

Fig. 16  (continued)

the experimental results of only three buildings (faculty and staff G + 2 housing) located in
the same campus. However, damping ratio using AVT may differ from the case of seismic
excitation.

7 Empirical estimate of natural time period and design


recommendations

Most seismic code recommendations are based on a regression analysis of the database
generated experimentally (or analytically) with height of the building as the key vari-
able. Table 1 summarizes the recommendations of Indian seismic standard along with
some of the international standards. Close scrutiny of Table  1 infers that the empiri-
cal estimate of natural period takes a general form of T = C ⋅ hk , where h is the overall
building height, k is an index and C is a coefficient. Parameters C and k are determined
through regression analysis but differ from one standard to another. Regardless of the
material (steel or concrete), most seismic codes consider k = 0.75 . This value is based

13
Bulletin of Earthquake Engineering

Fig. 17  Histogram constructed from the sample space of estimated damping ratio

on a set of assumptions including (1) a triangular variation of lateral forces over the
building height, (2) base shear proportional to T −𝛾  , with 𝛾 = 2∕3 , (3) uniform weight
over the height, and (4) uniform inter-storey drift due to applied lateral force profile.
The parameter C  , on the other hand, depends on the material type, for example, steel
or concrete. The equation is applicable specifically to the frame buildings. Most seis-
mic standards also recommend a simple functional form, namely, T = 0.1N  , where N
denotes number of storeyes, in particular for low-rise buildings (less than twelve sto-
ries). Empirical estimate of natural period for buildings with shear walls (concrete or
masonry) also take a similar form but with k = 1.0 . However, the parameter C is usually
expressed as a function of the effective cross-sectional area of shear walls placed in
the horizontal direction under consideration and the height of building. Two different
types of recommendations are apparent. The recommendation of ASCE 7-16 (2017) is
of different type when compared with that of UBC 1997, New Zealand Standard NZS
1170-5-2004 and Eurocode 8. A detailed derivation of the ASCE 7-16 (2017) equation
is presented by Goel and Chopra (1997). The equation is based on statistical analysis
of natural periods of buildings obtained by analyzing data from instrumented build-
ings exposed to several Californian earthquakes. National Building Code of Canada
(NBCC 2015) recommends similar empirical expressions for natural periods regardless
of the presence of shear walls except that the coefficient C assumes a different numeri-
cal value. Indian Standard IS 1893 Part-I (2016) recommendation is somewhat like
UBC 1997 in case of buildings with concrete structural walls or unreinforced masonry
infill walls. However, in case of "other buildings" (including masonry infill wall pan-
els), IS 1893 Part-I (2016) assumes k  = 1, but the coefficient C is a function of the plan
dimension in the direction under consideration (against a constant value for the framed
buildings). Empirical expressions in most seismic codes are intentionally calibrated to
underestimate the true natural period and thereby overestimating the design base shear.
The order of underestimation in natural period usually increases in the regions of severe

13
Bulletin of Earthquake Engineering

seismicity. While this factor varies from1.4 to 1.7 in ASCE 7-16 (2017), 1893-Part I
(2016) does not recognize such distinction.
Therefore, the empirical estimates recommended by most seismic standards including
the Indian Standards for similar type of RC buildings without structural walls but including
masonry infill panels (and tacitly extrapolated to the confined masonry buildings by the
practicing engineers in absence of appropriate design recommendation, which is question-
able) take the general form

T = 𝛼H 𝛽 Dk (19)
Indian standard recommends, 𝛼 = 0.09 , 𝛽 = 1 , and k = −0.5 where, H is overall height
of the building and D is the plan dimension along the direction of vibration. In case of
building with structural walls, the general recommendation is of the form
( )k
T = 𝛼H 𝛽 Aw (20)

with Aw as the total effective wall area. Indian Standards recommends Eq. (20) with
𝛼 = 0.075 , 𝛽 = 0.75 , and k = −0.5 subjected to an upper limit given by Eq. (19).

7.1 Assessment of Indian standard recommendations against experimental results

A total of eighteen natural periods are extracted in this paper experimentally, two for each
building. Torsionally dominated mode from the triplet of fundamental frequencies is not
considered here. Since all the buildings have same height, few more estimates are taken
from the testing of G + 2 housing units located in the same campus and more details of
which were reported in Nema and Basu (2019). This is intended to arrive at a relatively
diverse building stock and hence, a robust empirical estimate. Relevant geometric data
(total effective wall area (column 2), plan dimension along the direction of vibration (col-
umn 3), overall height (column 4) and number of stories (column 5)) of the entire building
stock along with the experimentally determined natural periods (column 6) are summa-
rized in Table 9.
In absence of any specific recommendation on confined masonry buildings, it seems to
be a critical decision making whether (1) Eq. (20) is appropriate assuming a case closer to
the building with RC structural walls subjected to Eq. (19) as an upper limit or (2) Eq. (19)
is suitable under the assumption of RC frame building with masonry infill walls. First, Eq.
(20) is used to for estimating the periods that are denoted as the estimate-A and presented
in column 7. Associated error with respect to the experimental period as the benchmark is
included in column 8 and the root-mean-square-error (RMSE) is noted as 41%. Second,
Eq. (19) is for the similar assessment and the associated results are shown in columns -9
and -10 with RMSE noted as 39%. Third, following the Indian Standard recommendation
for building with RC shear walls and using Eq. (20) with an upper limit given by Eq. (19),
results are included in columns -11 and -12. RMSE is noted as 21%. At this stage, it should
be noted that Eq. (20) for the interpretation of structural wall in RC frame building does
not lead to better estimate than Eq. (19) for the interpretation of masonry wall in RC frame
building. A conservative estimate (lower period) between two recommendations reduce the
error considerably, for example, RMSE is reduced 21% from 41%. In other words, the natu-
ral period of confined masonry buildings cannot be fully described by either structural wall
or infill wall interpretation as reflected by the comparison of RMSE in both the cases.

13
Table 9  Comparison of experimental and empirical estimates of fundamental frequency for the entire building stock
Building Effective Length, Height, No of Experi- IS 13920 Recommendations Proposed Recommendations
Wall D (m) H (m) Storey mental
Area, Aw Period Estimate-A Estimate-B Estimate- Min Estimate-1 Estimate-2
(m2) (sec) (A, B)
Period Error Period Error Period Error Period Error Period Error (%)
(sec) (%) (sec) (%) (sec) (%) (sec) (%) (sec)

(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16)
B_W1 6.28 10.1 12.8 4 0.24 0.20 15.6 0.36 51.1 0.20 15.6 0.24 0.1 0.24 1.7
Bulletin of Earthquake Engineering

B_W1 6.28 39.1 12.8 4 0.19 0.20 6.6 0.18 3.0 0.18 3.0 0.18 3.1 0.19 0.4
BW_2 2.77 10.1 12.8 4 0.25 0.30 21.9 0.36 45.1 0.30 21.9 0.24 3.9 0.24 2.4
BW_2 1.60 21.0 12.8 4 0.2 0.40 100.6 0.25 25.8 0.25 25.8 0.21 2.6 0.21 3.6
B_W3 3.93 10.1 12.8 4 0.25 0.26 2.4 0.36 45.1 0.26 2.4 0.24 3.9 0.24 2.4
B_W3 5.73 38.6 12.8 4 0.19 0.21 11.6 0.19 2.4 0.19 2.4 0.18 3.0 0.19 0.3
C_W1 6.51 11.1 12.8 4 0.24 0.20 17.1 0.35 44.1 0.20 17.1 0.23 2.1 0.24 0.8
C_W1 5.49 44.2 12.8 4 0.2 0.22 8.3 0.17 13.3 0.17 13.3 0.18 9.7 0.19 6.7
C_W2 2.22 10.1 12.8 4 0.24 0.34 41.9 0.36 51.1 0.34 41.9 0.24 0.1 0.24 1.7
C_W2 1.20 15.7 12.8 4 0.22 0.46 110.6 0.29 32.0 0.29 32.0 0.22 1.2 0.22 0.3
C_W4 3.88 10.1 12.8 4 0.25 0.26 3.1 0.36 45.1 0.26 3.1 0.24 3.9 0.24 2.4
C_W4 2.41 22.8 12.8 4 0.2 0.33 63.5 0.24 20.5 0.24 20.5 0.20 0.9 0.20 2.1
E_W4 9.48 10.1 12.8 4 0.25 0.16 34.1 0.36 45.1 0.16 34.1 0.24 3.9 0.24 2.4
E_W4 9.74 44.9 12.8 4 0.18 0.16 9.7 0.17 4.5 0.16 9.7 0.18 0.1 0.19 3.5
E_W5 4.58 11.1 12.8 4 0.24 0.24 1.2 0.35 44.1 0.24 1.2 0.23 2.1 0.24 0.8
E_W5 2.63 27.0 12.8 4 0.22 0.31 42.3 0.22 0.8 0.22 0.8 0.20 11.0 0.20 9.6
F_W2 5.51 10.1 12.8 4 0.23 0.22 6.0 0.36 57.7 0.22 6.0 0.24 4.4 0.24 6.1
F_W2 3.72 32.9 12.8 4 0.2 0.26 31.6 0.20 0.4 0.20 0.4 0.19 5.4 0.19 3.4
HB_30/2 4.09 13.0 10.3 3 0.18 0.21 18.5 0.26 42.6 0.21 18.5 0.20 11.5 0.19 3.6
HB_27/3 3.00 10.7 10.3 3 0.19 0.25 31.0 0.28 48.9 0.25 31.0 0.21 9.7 0.19 2.3

13

Table 9  (continued)
Building Effective Length, Height, No of Experi- IS 13920 Recommendations Proposed Recommendations
Wall D (m) H (m) Storey mental
Estimate-A Estimate-B Estimate- Min Estimate-1 Estimate-2

13
Area, Aw Period
(m2) (sec) (A, B)
Period Error Period Error Period Error Period Error Period Error (%)
(sec) (%) (sec) (%) (sec) (%) (sec) (%) (sec)

HB_19/2 2.82 9.5 10.3 3 0.18 0.26 42.7 0.30 67.1 0.26 42.7 0.21 18.7 0.20 11.0
HB_30/3 4.09 13.0 10.3 3 0.18 0.21 18.5 0.26 42.6 0.21 18.5 0.20 11.5 0.19 3.6
RMSE 41.0 RMSE 39.0 RMSE 21.2 RMSE 7.0 RMSE 4.3
(%) (%) (%) (%) (%)
Bulletin of Earthquake Engineering
Bulletin of Earthquake Engineering

Fig. 18  Correlation of period with geometric properties

A different approach is next adopted in order to identify the comparatively better inter-
pretation from Eqs. (19)�and
√ (20). Figure  18a presents the correlation of experimentally
�√
obtained period against 1 D whereas Fig. 18b presents the similar data against 1 Aw  .
Clearly, the former plot represents relatively better correlation and hence, Eq. (19) �√ is
expected to be a better candidate. Next, Fig. 18c presents the correlation against H D
�√
and resulting correlation is considerably improved when compared with 1 D . Fig-
�√
ure  18d presents the correlation against H 0.75
A and resulting correlation does not
�√ w
improve considerably when compared with 1 Aw  . In line with this comparison, it may
be inferred that Eq. (19) may be preferred Eq. (20) for the estimation of natural period of
confined masonry buildings.
However, the RMSE in the order of 40% does indicate an acceptable recommendation
and hence, further investigation is required as presented in the next section.

7.2 Proposed recommendation on natural period of confined masonry buildings

It is evident from Eq. (19) that the seismic code recommendation on natural period
approaches to zero when the height of the building tends to zero. However, while calibrat-
ing the empirical formula for finding out the coefficients through linear regression analysis
of the experimental data, the building stock hardly includes any samples with height close
to zero. In that sense, it is questionable to constrain the regression for neglecting the offset.
Two possible alternatives, namely the Estimates-1 and -2, are proposed in this paper
while arriving at the natural period of the confined masonry buildings from the experimen-
tal results as described in what follows.

13
Bulletin of Earthquake Engineering

7.2.1 Estimate‑1

First alternative considers an offset representing the natural period of one storey build-
ing while excluding height of the ground storey from the overall height of building.
Resulting empirical formula, denoted hereafter as the Estimate-1, takes the following
form:
� �
H−s
T=𝛼 √ +𝛽 (21)
D

here k = −0.5 conforming to the Indian Standard, s denotes the height of ground storey
and the offset 𝛽 is to account for the contribution from the ground storey. Since the storey
height is about 3.2  m in the present building stock, linear regression analysis is carried
out using s = 3 m and that results in 𝛼 = 0.037 and 𝛽 = 0.126 with R2 = 0.74 . Predictions
of natural period using Estimate-1 are compared in Table 9 (column 13) with the experi-
mental results (column 6). The associated error is included in column 14 and the RMSE is
noted as 7%. Therefore, Eq. (21), denoted as Estimate-1, explains the experimental results
much better than the recommendation of Indian standard.
In strict sense, Estimate-1 does not conform to the theory of structural dynamics and
resembles more like an intuitive approach. A more robust approach with some funda-
mental basis is presented next.

7.2.2 Estimate‑2

With reference to Fig.  19a, one N -storey building (with one DOF per floor level) is
assumed to be contributed from one (N − 1) - storey building and the ground storey for
the purpose of estimating only the fundamental period. Figure 19b illustrates the respec-
tive fundamental mode shapes and the symbol ⊕ denotes suitable combination. Let the
total mass of the (N − 1) - storey building be ms ; 𝜂 and Ks be the effective mass par-
ticipation and effective stiffness, respectively, in the fundamental mode. The associated
undamped stick model for the fundamental mode is (shown ) in Fig. 19c. Stick model of
the ground storey is also shown in Fig. 19c where Kg mg represents the stiffness (mass)
of the ground storey. Figure 19d combines the fundamental mode of the (N − 1) - storey
building with the ground storey to construct a 2-DOF system and fundamental period of
which may be considered as an approximation of that of the N -storey building.
Assuming the 2-DOF system as a shear building, which may be reasonable for the
CM constructions, the associated undamped free vibration problem may be expressed as
[ ]{ } [ ]{ }
Kg + Ks −Ks 𝜙g 2 mg 0 𝜙g
−Ks Ks 𝜙s
=𝜔
0 𝜂ms 𝜙s (22)

Rewriting the eigen problem as follows


{ } [ ]−1 [ ]{ }
1 𝜙g Kg + Ks −Ks mg 0 𝜙g
= (23)
𝜔2 𝜙s −Ks Ks 0 𝜂ms 𝜙s

the associated characteristic polynomial may be expressed as

13
Bulletin of Earthquake Engineering

N storey Building =

(N-1) storey Building Ground Storey


(a) Decoupling of N- storied building into two sub-assemblages

Fundamental Mode
N- storey Building

Fundamental Mode (N-1) Fundamental Mode


storey Building Ground storey

(b) Decoupling for the fundamental mode

(c) Decoupling into two stick models for the fundamental period

(d) 2-DOF stick model to capture the fundamental period

Fig. 19  Improving the estimation of natural period using a sub-assemblage based approach

13
Bulletin of Earthquake Engineering

mg 1 𝜂ms

Kg
− 𝜔2 Kg

det ⎢ mg 𝜂ms 𝜂ms 1
⎥=0 (24)

Kg Kg
+ Ks
− 𝜔2

⎣ ⎦

Now /defining fundamental frequency of the (N − 1)  - storey building through


𝜔2s = Ks 𝜂ms and frequency of the ground storey system through 𝜔2g = Kg mg , one may
/

show under the assumption of/ identical mass at all floor levels that
ms Kg = ms mg mg Kg = (N − 1) 𝜔2g . Therefore, Eq. (24) may be simplified to
/ ( / )( / )

1 1 𝜂(N−1)

𝜔2g
− 𝜔2 𝜔2g

det ⎢ 1 𝜂(N−1) 1 1
⎥=0 (25)

𝜔2g 𝜔2g
+ 𝜔2s
− 𝜔2

⎣ ⎦

Let 1 𝜔21 and 1 𝜔22 be the solution of Eq. (25) and utilizing the first invariant property,
/ /

1 1 1 𝜂(N − 1) 1 𝜂(N − 1) + 1 1
2
+ 2 = 2 +
𝜔g 2
𝜔g
+ 2 =
𝜔s 𝜔2g
+ 2
𝜔s (26)
𝜔1 𝜔2

In terms of the time period,

T12 + T22 = [𝜂(N − 1) + 1]Tg2 + Ts2 (27)

here T1 is the fundamental period of the 2-DOF system and can be approximated as that of
the N  - storey building, T  ; T2 is the higher mode period of the 2-DOF system; Tg is the time
period of the ground storey as single DOF; and Ts is the fundamental period of the (N − 1)
storey building. Clearly, following the Indian Standard per Eq. (19),
� �
H−s
Ts = 𝛼 √ (28)
D

and realizing that T2 << T1 , and hence, T1 ≈ T  , one may approximate
1
�2
⎤2



H s
T = ⎢𝛼 2 √ + {𝜂(N − 1) + 1}𝛾 2 ⎥ (29)

⎣ D ⎥

here 𝛼 2 and 𝛾 2 are the constants to be ascertained though regression analysis and N is the
total number of stories while 𝜂 = 0.8 is recommended to account for the effective mass par-
ticipation in fundamental mode.
Linear regression analysis of the present building stock leads to 𝛼 = 0.055 and
𝛾 = 0.093 with R2 = 0.89 . It is interesting to note that 𝛾 = 0.093 obtained from the
regression analysis also reasonably represents the natural period of a single storey
building. Comparison of natural periods using Estimate-2 (column 15) with experimen-
tal results (column 6) is presented in Table 9. The associated error is tabulated in col-
umn 16 and the RMSE is noted about 4%.
Clearly, even the Estimate-1 exhibits much better resemblance with the experimen-
tal results when compared with Indian Standard [Eq. (19) or even minimum of Eqs.
(19) and (20)]. However, Estimate-2 exhibits the best resembles with the experimental

13
Bulletin of Earthquake Engineering

results owing to its solid theoretical background. Therefore, this paper recommends Eq.
(29) for the estimation natural period of CM buildings.

7.2.3 Inferences

Also included in this comparison is the present recommendation of Indian Standard,


denoted here as the Estimate-3 (Table  9). Clearly, even Estimate -1 exhibits much better
resemblance with the experimental results when compared with Indian Standard (Esti-
mate-3). However, Estimate-2 exhibits the best resembles with the experimental results
owing to its solid theoretical background. Therefore, this paper recommends Eq. (29) for
the estimation natural period of CM buildings.

8 Base shear distribution and design recommendations

Most seismic standards recommend the distribution of base shear along the height in pro-
portion to the inertial force when calculating the lateral( ) load profile for equivalent static
analysis. In general, inertial force at any floor level Fi is assumed to be proportional to
wi hki , where, wi is the seismic weight and hi is height of the storey above the ground. ASCE
7-16 (2017) recommends k = 2 for natural period exceeding 2.5 s, k = 1 for natural period
less than 0.5 s and linear interpolation for the intermediate range. Eurocode 8 (Part-1) rec-
ommends either k = 1 or the use of true first mode shape instead of hki . New Zealand Stand-
ard (NZS 1170 2004) recommends accounting for the higher mode effects by applying 8%
of the base shear at the top while distribution of the remaining 92% in proportion to the
inertial force with k = 1 . National Building Code of Canada (NBCC-Vol-1 2015) also rec-
ommends k = 1 where higher mode effect is not expected and otherwise, certain percentage
of the base shear at the roof level while remaining in proportion to the inertial force. Indian
Standard (IS 1893- Part-1 2016) recommends k = 2 wherever equivalent static method is
applicable. None of these standards explicitly makes recommendations on the design of
CM buildings.
ASCE 7-16 (2017) recommends a reduction of k as the natural period reduces and this
is to account for the resulting change in inertial force distribution. In a similar line, CM
buildings are expected to deform predominantly in a shear mode and thereby requiring a
careful assessment on the height wise distribution of the base shear. Similar study is not
yet reported in the prior art.
Numerical model duly fine-tuned of each of the 9 buildings are considered one by one.
A set of 21 seismic events is randomly selected from the PEER NGA West2 database.
One of the recorded horizontal components from each event is chosen for constructing a
ground motion bin of size 21 and detailed description of which is presented in Table 10.
Each numerical model is analysed against all the selected accelerograms but one at a time.
Any of the selected accelerograms is applied along both the lateral orthogonal directions
but one at a time. Therefore, the entire analysis results lead to an ensemble of a typical
response parameter with the sample size of 42.
Consider one typical linear elastic time history analysis with 5% modal damping. At
any particular instant, shear force at any floor level is computed by summing up the storey
shear up to that storey level from the top. Time series of this shear force is extracted and
absolute maxima is noted regardless of the time instant. Let this shear force be denoted as
Vi at the ith storey level. Note that V1 also defines the base shear from the dynamic analysis.

13
Bulletin of Earthquake Engineering

Table 10  Description of selected ground motions


RSN Earthquake Name Magnitude Epicentral Distance VS30 (m/s) ROT Recorded com-
D50— ponent
PGA (g)

1 ’Helena, Mon- 6 6.31 593.35 0.16 Horizontal 2


tana-01’
3 ’Humbolt Bay’ 5.8 73.49 219.31 0.04 Horizontal 2
6 ’Imperial Val- 6.95 12.98 213.44 0.23 Horizontal 1
ley-02’
11 ’Northwest Calif- 5.8 55.96 219.31 0.11 Horizontal 2
03’
12 ’Kern County’ 7.36 118.26 316.46 0.05 Horizontal 2
16 ’Northern Calif-02’ 5.2 43.83 219.31 0.07 Horizontal 1
23 ’San Francisco’ 5.28 11.13 874.72 0.09 Horizontal 2
25 ’Northern Calif-04’ 5.7 58.78 219.31 0.07 Horizontal 2
28 ’Parkfield’ 6.19 36.18 408.93 0.06 Horizontal 1
34 ’Northern Calif-05’ 5.6 29.73 219.31 0.18 Horizontal 1
36 ’Borrego Mtn’ 6.63 70.75 213.44 0.09 Horizontal 2
41 ’Lytle Creek’ 5.33 106.33 450.28 0.02 Horizontal 1
51 ’San Fernando’ 6.61 71.14 280.56 0.03 Horizontal 1
95 ’Managua, Nicara- 6.24 5.68 288.77 0.35 Horizontal 1
gua-01’
97 ’Point Mugu’ 5.65 18.1 248.98 0.10 Horizontal 2
98 ’Hollister-03’ 5.14 11.09 1428.14 0.11 Horizontal 1
101 ’Northern Calif-07’ 5.2 30.54 567.78 0.16 Horizontal 2
106 ’Oroville-01’ 5.89 12.58 680.37 0.08 Horizontal 2
121 ’Friuli, Italy-01’ 6.5 55.62 496.46 0.03 Horizontal 2
126 ’Gazli, USSR’ 6.8 12.81 259.59 0.70 Horizontal 1
130 ’Friuli, Italy-02’ 5.91 17.12 310.68 0.11 Horizontal 2

Let 𝛽̂i = Vi V1 defines the fraction of base shear as the shear (not the storey shear) force at
/

the ith storey level.


The generic expression for the design lateral load recommended by most seismic stand-
ards is given by

wq hkq
Fq = ∑N V
k 1
(30)
j=1 w j h j

here Fq is the lateral load at q floor level with wq as the associated seismic weight and hq
as the height above the ground; N is the number of floors; V1 is the design base shear and k
is a specified exponent. Utilizing Eq. (30), one may write
∑N
w hk
j=i j j
𝛽i = ∑N (31)
w hk
j=1 j j

13
Bulletin of Earthquake Engineering

Fig. 20  Comparison of Base Shear Distribution (BSD) ratio along the height in numerical model (hollow
circle) of the building B_W1 against all 42 ground motion and proposed distribution (asterisk)

Therefore, 𝛽̂i = Vi V1 computed from the numerical model may be considered as an


/

estimate of 𝛽i from Eq. (31). Hence, given the numerical model of a building and a particu-
lar recorded accelerograms, the exponent k of Eq. (31) may be ascertained in a least square
sense.
The process is repeated for all the buildings and all the ground motions applied along
both the orthogonal directions but one at a time. Resulting sample space / of k leads to a
mean of 0.42 with a standard deviation of 0.16. The estimated 𝛽̂i = Vi V1 is compared
with 𝛽i from Eq. (31) using k  = 0.42. The sample comparison is presented for the build-
ing B_W1 in Fig. 20 that includes all the 42 cases of time history analyses. Regardless

13
Bulletin of Earthquake Engineering

of the ground motion, type of buildings and analysis direction, the proposed way of
distributing base shear well conforms to the numerical results. Since numerical models
are already shown to be in compliance with the experimental modal characteristics, the
proposed base shear distribution may be considered reasonable for the seismic design
of CM buildings. Also one may compare these results with the associated seismic code
recommendations that are primarily aimed to address the RC and steel building stock:
k  = 0.42 arrived at in this paper for confined masonry buildings is much lower.

9 Conclusions

Ambient Vibration Testing (AVT) is carried out on a set of 9 confined masonry hostel
buildings (G + 3) recently constructed at the residential campus of Indian Institute of
Technology Gandhinagar. Recorded signatures are processed and modal characteristics,
namely, natural frequencies, mode shapes and damping ratio are extracted. Modal char-
acteristics are primarily restricted to the first triplet of fundamental modes. Each build-
ing is modelled numerically and fine-tuned followed by a comparison of natural fre-
quencies and mode shapes in numerical model and experimental results. The fine-tuned
numerical models are analysed against a set of 21 recoded ground motions arbitrarily
selected from the PEER NGA West2 database. Distribution of the base shear along the
height is investigated and compared with the recommendation by most seismic stand-
ards. Natural frequencies extracted experimentally (AVT) are also compared with the
empirical estimate recommended by Indian Standards for similar type of buildings.
Based on the limited investigation, this study makes recommendations of two different
types, as enumerated below.

9.1 Possible design recommendations on CM buildings

1. Empirical equation for the natural period of CM buildings is proposed and validated
based on the AVT conducted on similar constructions. Proposed recommendation
assumes the similar functional from recommended by most seismic standards for the RC
buildings with masonry infill walls duly modified. The modification considers several
factors including the number of storey and expected first mode participation. However,
the experimental stock comprises of the buildings of similar height (except a set of
three), constructed using similar masonry units and mortar, and founded over similar
soil types. Therefore, the functional form recommended in this paper of natural period
should be recalibrated using a variety of other building stocks.
2. Distribution of base shear along the height is proposed for CM buildings and validated
from the linear time history analysis of associated numerical models subjected to a
set of recorded ground motions. The numerical models considered for this validation
are tuned to the modal characteristics extracted using AVT. The proposed distribution
recommends the shear forces at any storey level rather than the lateral force at any floor
level with an exponent close to 0.4.
3. Median damping ratio of the building stock is noted around 5% of critical, which how-
ever, may not reflect the true scenario during a seismic excitation.

13
Bulletin of Earthquake Engineering

9.2 Recommendations on AVT framework

1. When compared with PoGER and PreGER, a slightly different merging strategy is
adapted in this paper does not distinguish between the setups and consequently, accounts
for the cross-spectral density (CSD) between the setups. Comparison of different merg-
ing strategies is beyond the scope of this paper and will be reported elsewhere whenever
available.
2. The adapted AVT framework works on the response at degrees of freedom (DOF) rather
than the recorded signals directly. Even for an over-determined case, it is recommended
to extract the signature at the DOF (under the assumption of in-plane rigidity of the
floor diaphragms) before carrying out the frequency domain decomposition (FDD). In
that sense, the adapted framework imparts better clarity to the interpretation of results
when compared with other existing frameworks.
3. Extracted mode shape associated with a torsionally dominated mode may not be accurate
enough. Mass information is required for employing the necessary corrections.
4. Presence of secondary structures like staircase may slightly affect the fundamental
period of the primary structure and introduce additional frequencies to the overall struc-
ture, which can be misleading in some sense. These frequencies must be interpreted
based on conceptual understanding.
5. Lateral-torsional coupling usually results in closely spaced modes within a triplet. A
GUI based approach utilizing the modal coherence is proposed for extracting/identifying
the natural frequencies.

Acknowledgements  This research is funded by the Council of Scientific and Industrial Research, India,
under the Grant No. 22/762/17/ EMR-II. Financial support is gratefully acknowledged. Authors also
acknowledge Mr. Abhi Mittal for his assistance while testing some of the buildings and Mr. Falak Vats dur-
ing preparation of the manuscript.

Author contributions  DB proposed the idea, interpret the results, prepared the final draft, acquire the fund-
ing, interpreted results, and manage the overall research. PC performed the experiments, processed the data,
analysed with custom code, and licenced software, generated results, and prepared the first draft.

Funding  This research is funded by Council of Scientific and Industrial Research, India, under the Grant
No. 22/762/17/ EMR-II and the financial support is acknowledged.

Availability of data and material  Data may be available from the corresponding author through making rea-
sonable request.

Code availability  Standard licenced software is used. Custom code is developed in MATLAB environment
and not available for sharing.

Compliance with ethical standards 


Conflict of interest  Authors declare that the manuscript is free from any Conflicts of interest and Competing
interests.

13
Bulletin of Earthquake Engineering

References
ASCE/SEI 7-16 (2017) Minimum design loads for buildings and other structures, Standard ASCE/SEI7-
16. American Society of Civil Engineers
Astroza M, Moroni O, Brzev S, Tanner J (2012) Seismic performance of engineered masonry buildings in
the 2010 Maule earthquake. Earthq Spectra 28(S1):S385–S406
Au S-K, Zhang F-L (2012) Ambient modal identification of a primary–secondary structure by Fast Bayesian
FFT method. Mech Syst Signal Process 28:280–296
Basu D (2019) Back-of-the-Envelope analysis of plan-asymmetric confined masonry buildings for force-
based seismic design. J Build Eng 21:455–467
Basu D, Whittaker AS, Constantinou MC (2017) On the design of a dense array to extract rotational compo-
nents of earthquake ground motion. Bull Earthq Eng 15(3):827–860
Bashir A, Basu D (2018) Revisiting probabilistic seismic hazard analysis of gujarat: an assessment of indian
design spectra. Nat Hazards 91(3):1127–1164
Bendat JS, Piersol AG (2011) Random data: analysis and measurement procedures, vol 729. Wiley,
Hoboken
Brincker R, Andersen P, Jacobsen N-J (2007) Automated frequency domain decomposition for operational
modal analysis. In: Proceedings of The 25th international modal analysis conference (IMAC), Orlando,
Florida. vol 415
Brincker R, Ventura C (2015) Introduction to operational modal analysis. Wiley, Hoboken
Brincker R, Zhang L, Andersen P (2000) Modal identification from ambient responses using frequency
domain decomposition. In: Proceedings of the 18th international modal analysis conference (IMAC),
San Antonio, Texas
Brincker R, Zhang L, Andersen P (2001) Modal identification of output-only systems using frequency
domain decomposition. Smart Mater Struct 10(3):441
Brownjohn JMW, Magalhaes F, Caetano E, Cunha A (2010) Ambient vibration re-testing and operational
modal analysis of the Humber Bridge. Eng Struct 32(8):2003–2018
Chakra-Varthy P (2019) Confined masonry buildings: evaluation of modal properties using ambient vibra-
tion testing. MTech Thesis, Indian Institute of Technology Gandhinagar
Conte JP, He X, Moaveni B, Masri SF, Caffrey JP, Wahbeh M, Tasbihgoo F, Whang DH, Elgamal A (2008)
Dynamic testing of Alfred Zampa memorial bridge. J Struct Eng ASCE 134(6):1006–1015
De Sortis A, Antonacci E, Vestroni F (2005) Dynamic identification of a masonry building using forced
vibration tests. Eng Struct 27(2):155–165
Eurocode-8 (2009) Design of structures for earthquake resistance: part 1: general rules, seismic actions and
rules for buildings. In: European Committee for Standardization, B-1050, Brussels
Felber AJ (1994) Development of a hybrid bridge evaluation system. PhD thesis. University of British
Columbia
Foraboschi P (2019) Masonry does not limit itself to only one structural material: Interlocked masonry ver-
sus cohesive masonry. J Build Eng 100831
Foraboschi P (2016) The central role played by structural design in enabling the construction of buildings
that advanced and revolutionized architecture. Constr Build Mater 114:956–976
Goel RK, Chopra AK (1997) Vibration properties of buildings determined from recorded earthquake
motions. Report No. UCB/EERC-97/14, Earthquake Engineering Research Center, University of
California
Goulet J-A, Michel C, Smith IFC (2013) Hybrid probabilities and error-domain structural identification
using ambient vibration monitoring. Mech Syst Signal Process 37(1–2):199–212
Heylen W, Lammens S, Sas P (1995) Modal analysis theory and testing, Department of Mechanical Engi-
neering, Katholieke Universiteit Leuven, Leuven, Belgium
Humar JL (2012) Dynamics of structures. CRC Press, Boca Raton
IS 1893-Part 1 (2016) Criteria for earthquake resistant design of structures Part-1: General Provisions and
Buildings, Bureau of Indian Standards, Manak Bhavan, New Delhi
IS 456 (2000) Plain and Reinforced Concrete—Code of Practice, Bureau of Indian Standards, Manak
Bhavan, New Delhi
Jacobsen N-J (2006) Separating structural modes and harmonic components in operational modal analysis.
In: Proceedings IMAC XXIV conference
Liu Y-C, Loh C-H, Ni Y-Q (2013) Stochastic subspace identification for output-only modal analysis:
application to super high-rise tower under abnormal loading condition. Earthq Eng Struct Dynam
42(4):477–498
Magalhaes F, Caetano E, Cunha A, Flamand O, Grillaud G (2012) Am-bient and free vibration tests of the
Millau Viaduct: evaluation of alternative processing strategies. Eng Struct 45:372–384

13
Bulletin of Earthquake Engineering

Meli R, Brzev S, Astroza M, Boen T, Crisafulli F, Dai J, Farsi M, Hart T, Mebarki A, Moghadam AS et al.
(2011) Seismic design guide for low-rise confined masonry buildings, Oakland, California 94612-
1934, Earthquake Engineering Research Institute
NBCC (2015) National building code of Canada. In: National Research Council and others, Ottawa, Canada
Nema H, Basu D (2019) Natural properties of confined masonry buildings–experimental case studies and
possible inferences. Int J Masonry Res Innov 4(3):197–226
NZS 1170-5 (2004) Structural Design Actions, Part 5: Earthquake Actions—New Zealand
Parloo E, Guillaume P, Cauberghe B (2003) Maximum likelihood identification of non-stationary opera-
tional data. J Sound Vib 268(5):971–991
Parloo E (2003) Application of frequency-domain system identification techniques in the field of operational
modal analysis. Vrije Universiteit Brussel, Belgium
Peeters B, De Roeck G (1999) Reference-based stochastic subspace identification for output-only modal
analysis. Mech Syst Signal Process 13(6):855–878
Peeters B, De Roeck G (2001) Stochastic system identification for operational modal analysis: a review. J
Dyn Syst Meas Contr 123(4):659–667
Rainieri C, Fabbrocino G (2014) Operational modal analysis of civil engineering structures: An introduction
and a guide for applications. Springer, New York
Reynders E, Magalhaes F, De Roeck G, Cunha A (2009) Merging strategies for multi-setup operational
modal analysis: application to the Luiz I steel arch bridge. In: Proceedings of IMAC. vol 27
Rodda GK, Basu DB (2020) A novel framework for conditional simulation of fully nonstationary spatially
varying ground motion field. Earthq Eng Struct Dyn. https​://doi.org/10.1002/eqe.3343
Trifunac MD (1972) Comparisons between ambient and forced vibration experiments. Earthq Eng Struct
Dyn 1(2):133–150
Turek M, Thibert K, Ventura C, Kuan S (2006) Ambient vibration testing of three unreinforced brick
masonry buildings in Vancouver, Canada. In: Proceedings of the 24th International Modal Analysis
Conference (IMAC), Saint Louis, MI, USA
UBC (1997) Uniform building code. In: International Conference of Building Officials, Whittier, CA
Valente M, Milani G (2019) Damage assessment and collapse investigation of three historical masonry pal-
aces under seismic actions. Eng Fail Anal 98:10–37
Valente M, Milani G (2018) Damage assessment and partial failure mechanisms activation of historical
masonry churches under seismic actions: three case studies in Mantua. Eng Fail Anal 92:495–519
Verma R, Basu D (2017) On correlating the modulus of elasticity of stack- bonded flyash brick masonry
using impact hammer and compression tests. Eur J Civ Environ Eng. https​://doi.org/10.1080/19648​
189.2017.14102​32
Vestroni F, Beolchini GC, Antonacci E, Modena C (1996) Identification of dynamic characteristics of
masonry buildings from forced vibration tests. In: Proceedings of the 11th world conference on earth-
quake engineering
Xie Y, Liu P, Cai G-P (2016) Modal parameter identification of flexible spacecraft using the covariance-
driven stochastic subspace identification (SSI-COV) method. Acta Mech Sin 32(4):710–719
Zhang L, Brincker R (2005) An overview of operational modal analysis: major development and issues. In:
1st international operational modal analysis conference. Aalborg Universitet, pp 179–190

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

13

You might also like