You are on page 1of 18

Mechanical Systems and Signal Processing 151 (2021) 107396

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


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

Quantifying local stiffness loss in beams using rotation rate


sensors
L. Huras, Z. Zembaty ⇑, P.A. Bońkowski, P. Bobra
Faculty of Civil Engineering and Architecture, Opole University of Technology, Poland

a r t i c l e i n f o a b s t r a c t

Article history: The application of rotation rate sensors in vibration-based damage detection is investi-
Received 10 May 2020 gated. For this purpose, simulation, FEM analyses of a damaged cantilever plexiglass beam
Received in revised form 21 October 2020 under kinematic excitations are carried out and verified experimentally. Three types of
Accepted 25 October 2020
rotation rate sensors are examined. Damage index based on the maximum difference of
Available online 14 November 2020
signals on both sides of the damaged location is compared with averaged damage indices
in time and frequency domain. It is demonstrated that the rotation velocities of the beam
Keywords:
axis are less sensitive to measurement noise than the transversal accelerations. A simple
Vibrations
Beam flexure
damage index based on peak response and two, integral damage indices were tested using
Damage detection numerical simulations and laboratory experiments. The integral time-domain damage
Rotation rate sensors index appeared to be equally useful in damage detection, like the index based on the trans-
Time-history response missibility concept. The results of the experiments show positive prospects for construct-
ing Structural Health Monitoring (SHM) networks using rotation rate sensors. This is
particularly important for future SHM systems to be commonly implemented in civil
engineering.
Ó 2020 Elsevier Ltd. All rights reserved.

1. Introduction

Sensors measuring angular velocity (rotation rate sensors) find more and more applications in research of geophysics
(e.g., [17,22]), vehicle [33], aircraft motion control, e.g., drones [29], spacecraft engineering [6], human posture monitoring
[30] or even consumer electronics (e.g., smartphones or smartwatches, [4]). A critical analysis of selected rotation rate sen-
sors can be found in a review paper by Jaroszewicz et al. [18].
Rotational degrees of freedom are studied in the case of complex structural systems requiring sub-structuring [5] or are
used together with translations in 6-dof excitations of tested structural systems [15]. Measuring the rotational field of vibrat-
ing beams can substantially improve modern, wavelet-based SHM methods (e.g., [19]).
Thus, a specific need for measuring rotations of some elements and parts of structures occurs (e.g. [7]). The rotation rate
sensors can find particular application in Structural Health Monitoring (SHM) of civil engineering structures. The implemen-
tation of vibration-based damage detection algorithms for these types of structures is facing difficulties due to the complex-
ity of such the building materials like reinforced concrete or masonry and when including soil-structure interaction effects
(see, e.g., [27]). However, the specific abilities of the rotation rate sensors are not yet adequately appreciated in SHM of build-
ings and civil infrastructure (see e.g., a review paper by [11]). This is especially important in the light of the reported

⇑ Corresponding author.
E-mail address: z.zembaty@po.edu.pl (Z. Zembaty).

https://doi.org/10.1016/j.ymssp.2020.107396
0888-3270/Ó 2020 Elsevier Ltd. All rights reserved.
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

difficulties of the practical applications of transmissibility concept in vibration-based damage detection, as indicated in the
paper by Chesne and Deraemaeker [8].
On the other hand, particular advantages of these sensors in improving vibration-based damage detection in civil engi-
neering have already been theoretically and numerically confirmed (e.g., [1,21]). Besides, some laboratory tests in small
scales were carried out, too [2–3,37–38]. Very recent paper by Huseynov et al. [16] provide further successful results of
the application of rotational measurements in SHM of bridge structures.
The rotation rate sensors can be used in SHM in various ways, as illustrated in Fig. 1:

 Two rotation rate sensors may be used to monitor plastic hinge development in selected places of a frame structure under
seismic excitations (Fig. 1a). Warning of plastic hinge appearance may improve post-earthquake disaster mitigation or
provide early warning in case of a moving load on a bridge (Fig. 1b).
 Other two rotation rate sensors may serve to monitor curvature changes j = 1/q of beam axis during excessive vibrations
[37]. This is particularly important for partly cracked reinforced concrete beams when the effective application of strain
gauges is not possible (left side of Fig. 1b). The presence of inherent cracks in reinforced concrete structures during their
normal exploitation contributes to difficulties in their monitoring. These problems are studied for a long time in order to
observe the dynamic behavior of reinforced concrete structures (e.g., [23,39] and many more papers).
 A set of translational and rotational sensors (Fig. 1c) can be used for damage localization and quantification of a structure
under dynamic load [20–21,38].
 Even a single rotational sensor can effectively be used to provide interstory drift from axis rotation located in the middle
of a column, as shown in a paper by Schreiber et al. [31].

Fig. 1. Examples of the application of rotation rate sensors in SHM of civil engineering, a) moment resisting frame subjected to seismic excitations causing
plastic hinge formation in the frame corner, b) plastic hinge modes of failure of a beam and monitoring local curvature j = 1/q using two rotation rate
sensors, c) a slender tower in flexure, with bending stiffness EJ, under seismic excitations with an array of translational-rotational sensors and a damage in
its certain section.

2
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

Consider now the situation depicted in Fig. 1c. A question appears: To what extend by adding additional rotational mea-
surements to the existing transversal measurements, one can improve damage localization and detection? Such the question
was not formulated in the 90-ties of the XX century when the rotational sensors were expensive, often massive, and used
mostly in military applications e.g., in rocket motion control or satellite navigation. Nowadays, the decreasing price of elec-
tronic equipment makes it possible to upgrade existing SHM networks of 3-component translational accelerometers to 6-
component ones, including also angular velocities. This leads to 6 degree of freedom (6-dof) sensors in 3D or 3-dof sensors
in 2D, plane vibrations. Indeed, one may expect that including rotation rate sensors in dense arrays of SHM of structures
(Fig. 1c) may substantially improve localization and quantification of the damage. Rotation rate sensors were already applied
in vibration-based damage detection with positive results not only to acquire strain and local stiffness loss [37] but also in
more complicated, inverse tasks of stiffness reconstructions [38]. In their two recent papers, Al-Jailawi and Rahmatalla [2,3]
revived the transmissibility concept of SHM for damage detection using rotation rate sensors. Their experiments are, how-
ever limited to damages introducing local, flexural stiffness drops of 16 to 30%.
From the above literature review, it can be concluded that the advantages of rotation rate measurements in SHM are not
enough appreciated among researchers. There are very few numerical simulations published [1,20–21]. In particular, sensi-
tivity of the rotation rate measurements during laboratory or in situ measurements are not confronted with the demands of
SHM. The purpose of this paper is to carry on research in this direction and to quantify damage detection using three types of
rotation rate sensors, with various accuracy and prices. Following the positive results of earlier experiments [37], the ques-
tion is raised how small damages can be diagnosed using two rotation rate sensors of various types. It should be noted that
this can also help to address the problem of the plastic hinge monitoring before expensive shaking table experiments are
organized for moment resisting frames (Fig. 1a). It is so because the localized damage, as depicted in Fig. 1c may serve as
a simulation of local stiffness drop, similar to the effect caused by a plastic hinge developed under heavy load (Fig. 1b) or
strong seismic excitations (Fig. 1a). Installing rotation rate sensors on both sides of a localized stiffness loss may help find
discontinuity in the 1st derivative of horizontal beam axis or in the corner of the moment-resisting frame, which may serve
in early warning of damage or plastic hinge formation. In the research reported in this paper, numerical finite element anal-
ysis and subsequent laboratory experiments demonstrate the particular usefulness of the rotation rate sensors in vibration-
based damage detection of as small damages as 3.65% local reduction of cross-section, effectively indicated for plexiglass
cantilever model beam. For this purpose, three damage indices are analyzed. It is also shown that the rotation rate sensors
are particularly effective in damage quantification using averaged indices not only in the frequency domain but also with a
new, proposed damage index in the time domain.

2. Problem statement and finite element model formulation

Consider two examples of large, engineering structures equipped with networks of translational-rotational sensors
(Fig. 2):

 A simply supported beam representing, in a simplified way, a bridge structure

and

 a portal frame.

In the case of the bridge structure, one can excite it using a special, dedicated vehicle accessing the bridge when it is still
in an intact state and later, when a possible local stiffness loss occurs in a particular section of the bridge. The vehicle passing
the bridge with specific velocity is able to remotely acquire data from the sensors. In the second case, one can consider a
portal frame before an earthquake and afterwards. In this case, a specially designed harmonic shaker can be applied to excite
vibrations, which will be acquired from respective sensors close to the alleged ‘‘damage” location. The harmonic shaker (ex-
citer) can generate ordinary harmonic excitations or any signal specially dedicated for SHM purposes, e.g., a modulated pulse
of harmonic excitations (Fig. 3).
This way, rotational-translational responses can be acquired under the same excitations, yet with differences introduced
by the local stiffness loss. However, to carry on simple tests of the proposed methodology in a laboratory, one needs to design
a simplified experiment. Such a simple experiment may use a model of a cantilever beam under kinematic excitations, as was
already the case in previous research (e.g., [38]). In this case, one can use the kinematic excitations in the form of harmonic or
seismic excitations instead of the moving load or the exciter mounted to the structure. Following this approach, to be later
verified in the laboratory, consider now a small fragment of a cantilever beam (Fig. 4). The beam has only 4 degrees of free-
dom measured: two transversal qi(t), qi+1(t) and two rotational ui(t), ui+1(t), and a local loss of stiffness between them. The
horizontal position of the beam from Fig. 1c is motivated by the planned experiment, in which the kinematic excitations will
be applied in vertical direction. The intention of this simplification is to explore the possibility of:

a) assessing damage between two sensors of the array of Fig. 1c using only the measurements from the adjacent sensors
numbered ‘i’ and ‘i + 10 ,

3
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

Fig. 2. Simple implementation of rotational sensors in practical SHM applications.

Fig. 3. Acceleration record of modulated harmonic ‘pulse’ of excitations, and its Fourier transform.

Fig. 4. Cantilever beam under vertical kinematic excitations u(t) with local loss of cross section area (drop in flexural stiffness) and adjacent sensors
€iþ1 and angular velocities u
€i , q
measuring translational accelerations q _ i, u
_ iþ1 .

4
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

b) quantifying local stiffness loss due to plastic hinge development (Fig. 1a and 1b). It is interesting to note that this is
equivalent with measuring a discontinuity in the first derivative of the beam axis.

The problem of localizing the segment (or segments) with stiffness losses goes beyond the scope of the research reported
in this paper.
The beam shown in Fig. 4 will be excited by vertical, kinematic movement u(t) of its support using steady, harmonic exci-
tations and the harmonic ‘pulse’ from Fig. 3. In addition the familiar, seismic acceleration record El Centro [10,13] repre-
sented by the starting 20 s of the N-S component of this record (Fig. 3) will be applied too. The El Centro record is
frequently used as a benchmark in seismic engineering and provides clear wideband excitations useful in interpreting the
dynamic structural response. For the convenience of this analysis sampling frequency 50 Hz of the original record was trans-
formed into 1000 Hz using parabolic interpolation.
It can be seen from Fig. 5 that the power spectral density of this record is broad-band, dominating mostly in the frequency
range from 0 to 6 Hz.
The investigated cantilever beam 100 cm long, is assumed to be made of plexiglass, with cross-section 1.92 cm  11.5
0 cm and Young modulus E = 3.427 . 109N/m2 (density q = 1.19 g/cm3). The damping ratio applied in the computations
equaled 3.94% as obtained from free vibration tests of plexiglass beam applied in the experiments described in the reminder
of this paper. The beam is modeled using SAP 2000 program [9] and divided into 100 finite elements. As an effect of this
modeling, the distance Dx between the typical, adjacent sensors is divided onto 10 finite elements. The damage is included
in the middle of the beam (50 cm), and the response is measured at two coordinates: 45 and 55 cm from the beam support.
The damage is modeled as a loss of the cross-section area, leading to a decrease in bending stiffness. The damaged area
extends for 2 cm i.e. along 2 finite elements (Fig. 6). Results of the eigenproblem of the beam yield the first natural period
of the intact beam T1 = 0.19 s (f1 = 5.26 Hz).
The response characteristics to be measured are translational (transversal) accelerations q €ðtÞ and rotation rates u _ ðtÞ. It is
obvious, that even for the intact beam, the responses q €iþ1 ðtÞ as well as u
€i ðtÞ and q _ i ðtÞ and u
_ iþ1 ðtÞ will differ even at a short
distance Dx along the beam, between coordinates i and i + 1.
In Fig. 7, the difference between transversal accelerations (a) and rotational velocities (b) at the distances 45 cm and
55 cm from the support (Dx = 10 cm) are presented, as computed using SAP 2000 FEM program for the El Centro excitations
(Fig. 5). The problem to be investigated is how these differences will be vulnerable to the presence of the damage in the mid-
dle between the two points i and i + 1.
The simplest possible way to observe the effect of the presence of damage on these response differences is to compare
relative peak response. One can observe damage induced variations in the values of peaks of the acceleration differences:

Pdiff € €
trans ¼ qiþ1 ðt Þ  qi ðt Þ ð1Þ
as well as in peaks of differences in the rotation rates:

rot ¼ uiþ1 ðt Þ  ui ðt Þ
Pdiff _ _ ð2Þ
Thus, the first measure of damage can be formulated as:

Fig. 5. Acceleration record of the N-S component of El Centro, May 18th 1940 (first 20 s) and its power spectral density.

5
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

 
 Pdiff diff 
 damaged  Pintact 
DI1 ¼  
 ð3Þ
 P diff
intact 

where P diff diff


intact denotes peak of the response difference among sensors i and i + 1 of the intact structure and P damaged denotes the
same peak of the response difference for the damaged structure. The only problem is at what time to choose the response? As
a logical choice, looking for peaks, when the ratio of peak response difference P diff
damaged to respective intact response difference

Pdiff
intact takes the maximum value, seems the best option.
To further investigate the ability of rotation rate sensors in damage detection, three methods will be tested and
compared:

Fig. 6. Finite element mesh of the investigated beam with the locations of analyzed responses (45 and 55 cm from the support). Two elements of reduced
stiffness represent damage of the beam.

€iþ1 ðtÞ  q
Fig. 7. Response of the model of intact beam to vertical excitations by scaled El Centro N-S record. Difference between accelerations q €i ðtÞ (a) and
rotation rates u
_ iþ1 ðtÞ  u
_ i ðtÞ (b) for the responses at distances of 45 cm and 55 cm (Dx = 10 cm) from the support of the beam.

6
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

a) The maximum difference of angular velocity related to the same difference of vibrations of the intact structure (eq.
(3)).
b) The cumulated ratio of adjacent rotation rate responses on both sides of the damage calculated for the entire time-
history response of the structure and normalized with respect to intact response:

Pratio ratio
damaged  P intact
DI2 ¼ ð4Þ
Pratio
intact

where Pratio ratio


damaged and P intact are calculated during the response between time instances t1 and t2 , using the following equation
for the damaged and intact states, respectively:
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u 1 Z t2 u _ iþ1 ðt Þ 2

Pratio ¼u
tt  t dt ð5Þ
2 1 u_ i ðtÞ
t1

c) Transmissibility of the rotation rate responses integrated in the frequency domain, based on differences in the coher-
ency among the signals measured on both sides of the damage introduced to the beam.

The damage index based on transmissibility is derived below and in the Annex.
The concept of transmissibility was developed in the research of SHM at the end of XX century [34,36] and at the begin-
ning of XXI century (e.g. [25,28]) and applied with success in vibration-based damage detection (e.g. [24]). The transmissi-
bility depends on the structural response in the frequency domain in various response directions. It can be used as a measure
of changes in local structural stiffness caused by the damage. It can be formulated independently of excitations or (some-
times) as dependent on the excitations ([24], page 2477). The application of transmissibility in SHM was recently reviewed
by Yan et al. [40]. In this study, it is defined by simply dividing the analyzed structural responses along coordinates i and j, in
the frequency domain:

Q i ðxÞ
T ij ðxÞ ¼ ð6Þ
Q j ð xÞ

where Qi(x), Qj(x) stand for any structural response (e.g. translational displacements, velocities, accelerations or rotational
velocities etc.). It is shown in the Annex, that for plane vibrations of structures under kinematic (seismic) excitations, such
the transmissibility is, as expected, independent of the intensity of excitations.
For the purpose of vibration-based damage detection the transmissibility should be calculated in specific frequency range,
in which the real valued magnitude of coherence c
 
Sij ðxÞ2
cij ðxÞ ¼ ð7Þ
Sii ðxÞSjj ðxÞ

is high (see e.g. recent paper by Al-Jailawi, Rahmatalla [2]), where Sij(x) stands for co-spectral density of signals i and j, while
Sii(x), Sjj(x) are respective auto-spectral densities of the analyzed responses. Thus, the damage index for the rotation rate
responses u _ i ðtÞ, u
_ iþ1 ðtÞ measured by two adjacent sensors i and i + 1 (Fig. 4) and defined using transmissibility (eq. (6)),
should be computed only in selected frequency range. In this paper the transmissibility-based damage index is defined in
form of following integral:
Z x2 T damaged  T intact
dx
i;iþ1 i;iþ1
DI3 ¼ ð8Þ
x1 T intact
i;iþ1

in which respective transmissibilities are calculated for the beam in the damaged (Tdamaged) and intact (Tintact) states.

3. Numerical simulations

The numerical simulations are carried out for the 1-meter long cantilever beam described in the previous section, sepa-
rately for three types of kinematic excitations applied vertically:

 steady harmonic excitations with frequency f = 4.0 Hz


 a short time, narrowband excitation signal (Fig. 3) and
 wide-band El Centro N-S acceleration record (Fig. 5)

7
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

Table 1
Comparison of damage indices DI1, DI2, DI3 for three types excitations and seven damage levels (numerical simulations).

Loss of flexural DI1 [%] DI2 [%] DI3 [%]


stiffness of the cross-
Harmonic Modulated El Centro Harmonic Modulated El Centro Harmonic Modulated El Centro
section DEJ [%]
4.0 Hz harmonics N-S 4.0 Hz harmonics N-S 4.0 Hz harmonics N-S
component component component
0 (intact) 0 0 0 0 0 0 0 0 0
0.5 0.097 0.095 0.094 0.010 0.010 0.010 0.019 0.021 0.020
1 0.195 0.192 0.189 0.019 0.021 0.021 0.039 0.042 0.041
2.5 0.496 0.486 0.480 0.050 0.053 0.053 0.099 0.106 0.104
5 1.020 0.998 0.984 0.102 0.109 0.108 0.203 0.217 0.214
10 2.164 2.103 2.074 0.215 0.230 0.228 0.428 0.456 0.450
20 4.916 4.713 4.645 0.481 0.515 0.511 0.955 1.019 1.005
40 13.389 12.451 12.207 1.274 1.364 1.356 2.501 2.664 2.633

In Table 1 comparisons of the damage indicators DI1, DI2 and DI3 for the three excitations are shown for seven damage
levels 0.5% 1%, 2.5%, 5%, 10%, 20% and 40%.
All the results presented in Table 1 have very similar values, demonstrating that the selection of the excitation type has no
significant effect on three damage indices DI1, DI2, DI3.
The DI3 index was calculated for the frequency ranges

 x1 = 8p rad/s (f1 = 4.0 Hz) to x2 = 12p rad/s (f2 = 6.0 Hz) for El Centro excitations and
 x1 = 8p rad/s (f1 = 4.0 Hz) to x2 = 16p rad/s (f2 = 8.0 Hz) for harmonic excitations

This frequency range was decided based upon observations of the modulus of coherency (Eq. (7)) of two rotations for
coordinates i and i + 1. This is illustrated in Fig. 8, where the modulus of coherency stays above 0.99 in the frequency range
from zero up until 20 Hz, for the intact and damaged beam with 20% local loss of stiffness. The criterion chosen to define the

Fig. 8. The coherency of rotation rate signals (upper plot) and transmissibility of the two signals (lower plot) for intact and selected values of damages (El
Centro excitations).

8
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

above frequency windows is based on the value of coherency, which should be as close to 1.0 as possible. In the case of the
beam response to ‘‘El Centro” record, this bandwidth is narrower.
It can also be seen in Fig. 8 that also transmissibility keeps stable value until about 10 Hz.
One of the main issues in vibration-based damage detection is the presence of measurement noise, which can pollute the
actual experimental results and prevent successful applications of on-line alarm raising systems. Thus, for the purpose of this
study, the problem was numerically tested by adding Gaussian noise to the results of FEM computations.
In Fig. 9, the damage index DI1 (Eq. (3)) is presented with increasing damage level, as computed for rotation rate
responses and transversal accelerations, using the FEM beam model described earlier. These differences were computed
for the peaks of the responses in the time instance t = 2.55 s, where they appeared as the maximum relative differences.
The computations were carried out without added noise, as well as with 0.01%, 0.02%, and 0.2% RMS (root mean square)
noise. It can be seen from this figure that without the presence of noise, both types of response (translational and rotational)
properly reflect increasing damage measured by local stiffness drop. However, adding even a small amount of noise distorts
peak translational response and, to much less extend, distorts the rotational response.
This is in agreement with earlier experiments that demonstrated the advantages of rotation rate sensors over transla-
tional accelerometers in SHM [2,37].
In Fig. 10 results of the analysis of the effect of increasing noise contribution from 0 to 5% are presented for all three dam-
age indices DI1, DI2 and DI3 normalized with respect to their values without noise pollution shown as 100%. It can be seen
that the single value DI1 damage index is the fasted to lose its stability with the increasing noise (it reaches 240% for 2%
noise). Both indices, which are based on averaged responses in the time domain (DI2) and in the frequency domain (DI3),
respond much better to the measurement noise.

4. Laboratory experiments

4.1. Description of the experimental setup and testing methodology

The tested plexiglass cantilever beam is going to be vertically excited by a free supported, 5.5 m steel beam, excited by an
actuator resting in the middle of the steel beam (Fig. 11). The vibrations of the support of the cantilever plexiglass beam are

Fig. 9. Numerical simulations of relative difference in peak response (damage index DI1) with increasing damage, computed without and with
measurement noise with three levels: 0.01% RMS noise, 0.02% RMS noise and 0.2% RMS noise (El Centro excitations).

9
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

Fig. 10. The effect of noise on the results of simulations for the El Centro excitations and 20% damage.

Fig. 11. Photograph with a general view of the tested beam mounted to the steel, beam acting as a support with kinematic vertical motion.

controlled by Instron system of data acquisition and actuator control. The actual kinematic excitations of the plexiglass beam
are measured by vertical accelerometer. The tested area of the plexiglass beam is located in the middle of the beam with
three pairs of the rotational sensors put apart in a distance of Dx = 10 cm (Fig. 12) and additionally instrumented by two
vertical, translational accelerometers (PCB 3711E1110G). The rotation rate sensors applied in the experiment are:
p
 Gladiator G150Z-100–100 with range ± 100 deg/s, output noise (1r) 0.001 deg/sec/ Hz, bandwidth: 0 – 200 Hz, resolu-
tion <= 0.0005 deg/s, (sampling 1000 Hz)
p
 Horizon HZ1-100–100, with range ± 100 deg/s, output noise 0.025 deg/sec/ Hz, bandwidth: 0 – 60 Hz, resolution < 0.
004 deg/s, (sampling 1000 Hz)
p
 MPU 6050 with range ± 250 deg/s, output noise 0.005 deg/sec/ Hz @ 10 Hz, bandwidth: 5 – 256 Hz, resolution 0.008 –
0.061 deg/s, (sampling 456 Hz).

The Gladiator sensors are the most accurate of the three types mentioned above, so they are the primary ones for this
experiment. The Horizons are less accurate than Gladiators, while the MPU sensors are low-cost sensors used in simple
drones. So, both Horizons and MPU sensors were included to investigate how demanding are the damage detection indices
with respect to the quality of rotation rate measurements. In order to apply progressive damage in a convenient way, it was
decided to drill holes in a selected place, between the sensors (see Fig. 12). Drilling holes had an advantage over exchange of
beams with various ‘‘cuts” in the center because only one beam was used during the experiment with exactly the same stiff-
10
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

Fig. 12. Detail of the tested plexiglass beam with three pairs of Horizon sensors (on top of the beam), Gladiators (on the facing side of the beam), and MPU
6050 (hidden below the beam). The actual damage state of the beam is the smallest one (one hole with 4.2 mm diameter, 3.65% loss of flexural stiffness).

ness in the intact area, the same mass distribution and boundary conditions. During drilling the beam was fixed and verti-
cally mounted drill was used with a 4 mm drill diameter. The actual hole created in the beam, measured by caliper, reached
diameter / = 4.2 mm. The first, single hole reduced cross-section area by 3.65%. In order to properly keep the symmetry of
vibrations of the beam, the dynamic tests were carried only after each odd number of the drilled holes: 1, 3, 5, 7, 9, & 11
(Table 2). It should be pointed out that a hole smaller than 4 mm made it impossible for the beam response to differentiate
it from the response of the intact beam. When the number of holes reached 11, the dynamic analyses of the tests were
stopped achieving the reduction of cross-section area equal to 40.78% (Table 2). The application of holes to inflict ‘‘damage”
(local stiffness loss) to the beam had a disadvantage because it introduced an unclear overall flexural stiffness loss effect. To
bind experimental results with the numerical analysis of section2, an ‘‘equivalent” stiffness loss corresponding with each
number of holes was calculated. The equivalency was defined by achieving the same flexibility of the cantilever beam mod-
eled as a plate using thick, shell finite elements by SAP 2000 program. The FEM mesh was dense enough to properly cover the
holes. This FEM model is described in detail in the next sub-section. The flexibility of the FEM plate model and the corre-
sponding beam was measured by keeping the same tip displacement of the beam and plate model. The ‘‘equivalent” stiffness
loss of the beam was calculated for the ‘‘damaged” zone of 2 cm length (2 finite elements), exactly the same as applied in the
preceding numerical analysis. The reduction of local flexibility was applied by reducing the width along the 2 cm ‘‘damage”
zone (Fig. 6). The values of effective stiffness losses are given in the last column of Table 2. Considering typical experimental

Table 2
Selected phases of damage tests analyzed in this paper.

Test reduction of cross-section area Description ‘‘equivalent” local


phase flexural stiffness
loss
I 3.65% one hole / 0.62%
= 4.2 mm
II 10.96% three holes 1.95%
/ = 4.2 mm
III 18.26% five holes / 3.45%
= 4.2 mm
IV 25.57% seven holes 5.12%
/ = 4.2 mm
V 32.87% nine holes 7.03%
/ = 4.2 mm
VI 40.17% eleven holes 9.22%
/ = 4.2 mm

11
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

achievements of SHM the low limit level of detected damage (3.65% of the area of the whole cross-section) seems a good
prognostic for the application of rotation rate sensors in SHM when compared with other results (e.g., [12,32], or recent
study by Mosavi et al. [26]).
Details of the tested beam are exactly the same as of the beam analyzed in the numerical analysis of chapter two
(100 cm  1.92 cm  11.50 cm). The tests were carried out using modulated harmonic excitations in the vertical direction
with peak acceleration 0.8 m/s2 (Fig. 3) measured by the PCB 3711E1110G sensor, installed at the support of the tested plex-
iglass beam.
In Fig. 13 the rotation rate difference among the two Gladiator sensors on both sides of the damage is shown for the
response of intact beam and phase III damage (3.45% of equivalent stiffness loss). The differences between the response
of intact and damaged beam are clearly seen.

Fig. 13. A difference of rotation rate sensors on both sides of the damage location (Dx = 10 cm) shown for intact and phase III damage (5 holes, equivalent to
18.26% local stiffness drop) – upper plot. The lower plot shows the difference between the response of intact and the damaged beam. The applied vertical
excitations of the plexiglass beam model are shown in Fig. 3.

12
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

4.2. Results of the experiments

For each of the selected damage phases, additional FEM computations were carried out. In this case, however, a much
finer mesh was used to properly cover the type of damage inflicted on the beam by drilling holes modeled by dodecagons
with ‘shell’ elements, defined by SAP FEM software [35]. These types of finite elements are accurate for both thin or thick
plates or shells. The FEM mesh was constructed in such a way that the mesh for each considered number of holes was mod-
ified by removing necessary elements from the ‘‘full” intact FEM model. For example, in the case of five holes (18.26% loss of
cross-section area, 3.45% ‘‘equivalent” loss of stiffness), the number of finite elements equaled 10 846, and the respective
mesh is shown in Fig. 14.
In Table 3, the damage indices DI1, DI2, DI3 (Eqs. 3, 4, 8) are shown, as computed based on FEM results and acquired by
Gladiator sensors. These results are displayed in graphical form in Figs. 15–17. It can be seen from these figures that up to
about 10% stiffness drop the results of linear FEM computations are rather close to experimental results. For large local stiff-
ness drop of 30–40%, the numerical results underestimate the damage indices up to 20–30%. This can be explained by geo-
metric non-linear effects influencing vibrations of the locally damaged cantilever. It should also be noted that, as expected,
all the damage indices increase with the damage. This makes room for possible linearization of these results. In this case, one
should also investigate a wider variety of structural models.
It can be seen in Figs. 15–17 that all damage indices depart from the FEM models with increasing damage. This may be
explained by possible geometrical, non-linear effects increasing with substantial loss of concentrated stiffness drop. The
damage index DI2 seems slightly more uniform than the other two indices. It is also simpler than the damage index DI3 for-
mulated in the frequency domain and based on transmissibility (Eqs. 6–8 and Annex).
In Fig. 18 results of the damage index DI2 from the experiment are shown for the three tested sensors. The results of the
‘‘Horizon” sensor measurements do not deviate much from the ‘‘Gladiator” sensors showing their similar performance in this
experiment. The results obtained using MPU 6050 sensors are less uniform and deviate more from the FEM computations.
Thus, it will be more difficult to calibrate these measurements with respective increasing damage, though the prospects for
applying cheap rotation rate sensors for typical civil engineering structures seem very good. Any future designs of SHM net-
works incorporating rotation rate sensors will require particular calibration techniques adapted to the specific types of struc-
tures (e.g., reinforced concrete, masonry, etc.).

5. Discussion and conclusions

An analysis of the application of rotation rate sensors for damage detection is presented as an extension of the existing
SHM techniques. Three damage indices measuring the local stiffness loss are proposed and examined. It is observed that sim-
ple damage index DI1 based on a single peak response value is very vulnerable to noise contamination. Both integral damage
indices DI2 (in the time domain) and DI3 (in the frequency domain) appeared more stable under the noise pollution.
The numerical analysis was carried out by FEM analysis of a cantilever beam with 2 elements of reduced stiffness (out of
100). This analysis clearly indicated better fragility of rotation rate measurements to the variations of flexural stiffness in the
beam section between two rotational measurements (Fig. 9). The experiments were carried out using a single specimen of

Fig. 14. A finite element model of the plexiglass beam weakened by 5 holes.

13
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

Table 3
Comparison of damage indices DI1, DI2, DI3 for numerical simulations, and respective experimental values obtained using Gladiator sensors.

Phase of damage & ‘‘equivalent” loss of DI1 [%] DI2 [%] DI3 [%]
flexural stiffness DEJ
FEM Experiment FEM Experiment FEM Experiment
simulation (Gladiator) simulation (Gladiator) simulation (Gladiator)
I – one hole (0.62%) 0.200 0.480 0.020 0.032 0.041 0.053
II – three holes (1.95%) 0.618 0.718 0.063 0.112 0.126 0.258
III – five holes (3.45%) 1.039 1.852 0.106 0.188 0.212 0.359
IV – seven holes (5.12%) 1.465 2.374 0.150 0.268 0.298 0.526
V – nine holes (7.03%) 2.541 3.334 0.203 0.291 0.404 0.579
VI – eleven holes (9.22%) 2.541 3.3810 0.259 0.394 0.516 0.774

Fig. 15. Damage Index DI1 (eq. (3)) as a function of increasing damage levels from FEM simulations and measured by Gladiator sensors during the
experiment.

plexiglass beam and subsequent drillings of holes in the center of the beam. In order to compare the stiffness loss of the
beam by hole drilling with numerical analyses of rectangular ‘cuts’ of cross sections an ‘‘equivalent” local stiffness loss
was defined. The experiments demonstrated that rotation rate sensors are very efficient in vibration-based damage detec-
tion. In fact, this was achieved not by particular virtues of the transmissibility method, but rather it is the ‘‘damage” (local
stiffness loss) sensitivity of the adjacent rotation response measurements, which is the most important. The experiments
which were carried out made it possible to indicate ‘‘damage” effect as low as 3.65% of local reduction of the cross-

Fig. 16. Damage Index DI2 (eq. (4)) as a function of increasing damage levels from FEM simulations and measured by Gladiator sensors during the
experiment.

14
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

Fig. 17. Damage Index DI3 (eq. (8)) as a function of increasing damage levels from FEM simulations and measured by Gladiator sensors during the
experiment.

Fig. 18. Damage Index DI2 as a function of increasing damage levels from linear FEM simulations, and as measured by Gladiators, Horizons and MPU 6050
sensors.

section area. The older (and cheaper) Horizons have shown results very close to the results of the more expensive Gladiators.
The cheap rotation rate sensors MPU 6050 brought results more different from the linear FEM models with fine mesh. All the
damage indices have shown increasing deviation from linear FEM models due to increasing geometrical non-linear effects
caused by localized increasing stiffness drop (large deformations). The plots from Figs. 15–18 also show that even the ‘cheap’
rotation rate sensors can be effectively calibrated to indicate presence of damage by measuring loss of stiffness among the
sensors located on beams or frames. The results of investigations in quantifying localized stiffness loss in the straight beam
of this experiment lead to the conclusion that rotation rate sensors may also be applied in monitoring potential plastic hinge
development of moment resisting frames. The dependence between actual damage and the values of various damage indices
require experiments carried out for full-size structures and many rotation rate sensors connected within networks with
other vibration sensors. Further studies with this respect and calibrations with damages are required and are planned by
the authors of this paper.
Putting an array of 6-dof sensors along the monitored beams and frames as well as in their corners will result in more
efficient damage detection and quantification than just by using transversally oriented accelerometers. With increasing
accuracy and drop in prices, one is able to prepare dense arrays of 6-dof sensors in various configurations (Fig. 1c, Fig. 2),
which will substantially improve the SHM of civil engineering structures. The remaining problems requiring further, in-
depth research is location of damages for the frame structure with such dense networks of 6-dof sensors. This can be
achieved by applying modern methodology of optimization with the aid of genetic algorithms (e.g. [14,20]). Practical imple-

15
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

mentation of such developed methodology of SHM in civil engineering, will require more studies of various types of
structures.

CRediT authorship contribution statement

L. Huras: Conceptualization, Methodology, Software, Validation, Formal analysis, Investigation, Resources, Data curation,
Writing - review & editing, Supervision, Project administration. Z. Zembaty: Conceptualization, Methodology, Software, Val-
idation, Formal analysis, Investigation, Resources, Data curation, Writing - original draft, Writing - review & editing, Visual-
ization, Supervision, Project administration, Funding acquisition. P.A. Bońkowski: Conceptualization, Methodology,
Software, Validation, Formal analysis, Investigation, Resources, Data curation, Writing - review & editing, Supervision. P.
Bobra: Conceptualization, Methodology, Software, Validation, Formal analysis, Investigation, Resources, Data curation,
Supervision, Project administration.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

Acknowledgements

This research was partially supported by statutory funds of the Polish Ministry of Science and Higher Education NBS
12/19 and NBS 15/19. The authors also wish to thank dr Seweryn Kokot for help in numerical noise simulations and dr Bro-
nisław Je˛draszak for assistance in carrying out the experiments. The Authors would like also to thank the anonymous review-
ers for their insightful comments and suggestions which improved the paper.

Annex

The equation of motion of a structure vibrating under plane, kinematic excitations u(t) takes following form (Fig. 1c,
Fig. 4):

€ þ cq_ þ kq_ ¼ m1u


mq € ðtÞ ¼ f ðtÞ ða1Þ

where m, c, k are matrices of inertia, damping and stiffness respectively and symbol 1 stands for vector containing
values of 1:
1 ¼ f1; 1;    ; 1gT r with m ¼ diagfmg a diagonal matrix
Moving to frequency domain one can calculate structural response to steady state vibrations in the following form:

Q ðxÞ ¼ HðxÞFðxÞ ða2Þ

where H(x) = [Hik(x)] is frequency response matrix of the structure and F(x) is the Fourier transform of excitation force
vector f(t) from eq. a(1):

FðxÞ ¼ m1u
€ ðxÞ ða3Þ
In equation (a3) symbol u€ ðxÞ stands for Fourier transform of excitation accelerations u
€ ðtÞ.
For the simple, plane structure from Fig. 1c or Fig. 4 the vector of generalized responses q(t) (eg. displacements or rota-
tions) transformed to frequency domain equals
X
Q i ðxÞ ¼ u
€ ð xÞ mk Hik ðxÞ ða4Þ
k

Calculating ratio of Qi(x)/Qj(x) for two responses i and j leads to transmissibility

Q i ðxÞ
T ij ðxÞ ¼ ða5Þ
Q j ð xÞ

which, in this case, does not depend on intensity of excitations


P
Q i ðxÞ mk Hik ðxÞ
T ij ðxÞ ¼ ¼ Pk ða6Þ
Q j ð xÞ k mk H jk ðxÞ

as sometimes can be the case for simple definitions of ‘‘local” transmissibility ([24], page 2477).
16
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

References

[1] M.-A.-B. Abdo, M. Hori, A numerical study of structural damage detection using changes in the rotation of mode shapes, J. Sound Vib. 251 (2002) 227–
239, https://doi.org/10.1006/jsvi.2001.3989.
[2] S. Al-Jailawi, S. Rahmatalla, Transmissibility-based damage detection using angular velocity versus acceleration, J. Civ. Struct. Health Monit. 8 (2018)
649–659, https://doi.org/10.1007/s13349-018-0297-0.
[3] S. Al-Jailawi, S. Rahmatalla, Damage detection in structures using angular velocity, J. Civ. Struct. Health Monit. 7 (2017) 359–373, https://doi.org/
10.1007/s13349-017-0224-9.
[4] Ó. Belmonte-Fernández, A. Puertas-Cabedo, J. Torres-Sospedra, R. Montoliu-Colás, S. Trilles-Oliver, An indoor positioning system based on wearables
for ambient-assisted living, Sensors 17 (2017) 36, https://doi.org/10.3390/s17010036.
[5] T. Bregar, N. Holeček, G. Čepon, D.J. Rixen, M. Boltežar, Including directly measured rotations in the virtual point transformation, Mech. Syst. Sig.
Process. 141 (2020) 106440, https://doi.org/10.1016/j.ymssp.2019.106440.
[6] Challoner, A.D., Ge, H.H., Liu, J.Y., 2014. Boeing Disc Resonator Gyroscope, in: 2014 IEEE/ION Position, Location and Navigation Symposium - PLANS
2014. Presented at the 2014 IEEE/ION Position, Location and Navigation Symposium – PLANS 2014, pp. 504–514. https://doi.org/10.1109/
PLANS.2014.6851410
[7] C. Chen, T. Ma, H. Jin, Y. Wu, Z. Hou, F. Li, Torque and rotational speed sensor based on resistance and capacitive grating for rotational shaft of
mechanical systems, Mech. Syst. Sig. Process. 142 (2020) 106737, https://doi.org/10.1016/j.ymssp.2020.106737.
[8] S. Chesné, A. Deraemaeker, Damage localization using transmissibility functions: a critical review, Mech. Syst. Sig. Process. 38 (2013) 569–584, https://
doi.org/10.1016/j.ymssp.2013.01.020.
[9] Computers and Structures, Inc., 2019. SAP2000. Walnut Creek, California, USA.
[10] COSMOS: El Centro 1940-05-19 04:36:41 UTC [WWW Document], 1940. URL http://www.strongmotioncenter.org/vdc/scripts/event.plx?evt=88
(accessed 9.28.16).
[11] S. Das, P. Saha, A review of some advanced sensors used for health diagnosis of civil engineering structures, Measurement 129 (2018) 68–90, https://
doi.org/10.1016/j.measurement.2018.07.008.
[12] S.W. Doebling, C.R. Farar, M.B. Prime, A summary review of vibration-based damage identification methods, Shock and Vibration Digest 30 (1998) 91–
105.
[13] El Centro Earthquake [WWW Document] 1940 URL http://www.vibrationdata.com/elcentro.htm (accessed 9.28.16).
[14] Farhangdoust S., Tashakori S., Baghalian A., Mehrabi A., Tansel I., Prediction of damage location in composite plates using artificial neural network
modeling, Proc. SPIE 10970, Sensors and Smart Structures Technologies for Civil, Mechanical, and Aerospace Systems 2019, 109700I (27 March 2019);
doi: 10.1117/12.2517422.
[15] E. Habtour, R. Sridharan, A. Dasgupta, M. Robeson, S. Vantadori, Phase influence of combined rotational and transverse vibrations on the structural
response, Mech. Syst. Sig. Process. 100 (2018) 371–383, https://doi.org/10.1016/j.ymssp.2017.07.042.
[16] F. Huseynov, C. Kim, E.J. OBrien, J.M.W. Brownjohn, D. Hester, K.C. Chang, Bridge damage detection using rotation measurements – experimental
validation, Mech. Syst. Sig. Process. 135 (2020) 106380, https://doi.org/10.1016/j.ymssp.2019.106380.
[17] H. Igel, J. Brokesova, J. Evans, Z. Zembaty, Preface to the special issue on advances in rotational seismology: instrumentation, theory, observations and
engineering, J. Seismolog. 16 (2012) 571–572, https://doi.org/10.1007/s10950-012-9307-6.
[18] L.R. Jaroszewicz, A. Kurzych, Z. Krajewski, P. Marć, J.K. Kowalski, P. Bobra, Z. Zembaty, B. Sakowicz, R. Jankowski, Review of the usefulness of various
rotational seismometers with laboratory results of fibre-optic ones tested for engineering applications, Sensors 16 (2016) 2161, https://doi.org/
10.3390/s16122161.
[19] A. Katunin, H. Lopes, J.V.A. dos Santos, Identification of multiple damage using modal rotation obtained with shearography and undecimated wavelet
transform, Mech. Syst. Sig. Process. 116 (2019) 725–740, https://doi.org/10.1016/j.ymssp.2018.07.024.
[20] Kokot S., Zembaty Z., 2009a Damage reconstruction of 3d frames using genetic algorithms with Levenberg-Marquardt local search, Soil Dynamics and
Earthquake Engineering, vol.29, no2. February, pp.311-323.
[21] S. Kokot, Z. Zembaty, Vibration based stiffness reconstruction of beams and frames by observing their rotations under harmonic excitations —
numerical analysis, Eng. Struct. 31 (2009) 1581–1588, https://doi.org/10.1016/j.engstruct.2009.02.032.
[22] Lee, W.H.K., Celebi, M., Todorovska, M.I., Igel, H., 2009. Introduction to the Special Issue on Rotational Seismology and Engineering Applications.
Bulletin of the Seismological Society of America 99, 945–957. doi: 10.1785/0120080344.
[23] J. Maeck, G. De Roeck, Dynamic bending and torsion stiffness derivation from modal curvatures and torsion rates, J. Sound Vib. 225 (1999) 153–170.
[24] N.M.M. Maia, R.A.B. Almeida, A.P.V. Urgueira, R.P.C. Sampaio, Damage detection and quantification using transmissibility, Mech. Syst. Sig. Process. 25
(2011) 2475–2483, https://doi.org/10.1016/j.ymssp.2011.04.002.
[25] N.M.M. Maia, J.M.M. Silva, A.M.R. Ribeiro, The transmissibility concept in multi-degree-of-freedom systems, Mech. Syst. Sig. Process. 15 (2001) 129–
137, https://doi.org/10.1006/mssp.2000.1356.
[26] A.A. Mosavi, D. Dickey, R. Seracino, S. Rizkalla, Identifying damage locations under ambient vibrations utilizing vector autoregressive models and
Mahalanobis distances, Mech. Syst. Sig. Process. 26 (2012) 254–267.
[27] D.M. McCann, M.C. Forde, Review of NDT methods in the assessment of concrete and masonry structures, NDT and E Int. 34 (2001) 71–84, https://doi.
org/10.1016/S0963-8695(00)00032-3.
[28] A.M.R. Ribeiro, J.M.M. Silva, N.M.M. Maia, On the generalisation of the transmissibility concept, Mech. Syst. Sig. Process. 14 (2000) 29–35, https://doi.
org/10.1006/mssp.1999.1268.
[29] L.V. Santana, A.S. Brandão, M. Sarcinelli-Filho, Outdoor waypoint navigation with the AR.Drone quadrotor, in: 2015 International Conference on
Unmanned Aircraft Systems (ICUAS). Presented at the 2015 International Conference on Unmanned Aircraft Systems (ICUAS), pp. 303–311, 2015. doi:
10.1109/ICUAS.2015.7152304.
[30] T.M.O. Santos, M.F.S. Barroso, R.A. Ricco, E.G. Nepomuceno, É.L.F.C. Alvarenga, Á.C.O. Penoni, A.F. Santos, A low-cost wireless system of inertial sensors
to postural analysis during human movement, Measurement 148 (2019) 106933, https://doi.org/10.1016/j.measurement.2019.106933.
[31] K.U. Schreiber, A. Velikoseltsev, A.J. Carr, R. Franco-Anaya, The application of fiber optic gyroscopes for the measurement of rotations in structural
engineering, Bull. Seismol. Soc. Am. 99 (2009) 1207–1214.
[32] H. Sohn, C.F. Farrar, F.M. Hemez, D.D. Shunk, D.W. Stinemates, B.R. Nadler, et al., A review of structural health monitoring literature: 19962001. Report,
LA-13976-MS. Los Alamos (USA): Los Alamos National Laboratory, 2003.
[33] M.R. Solouk, M.H. Shojaeefard, M. Dahmardeh, Parametric topology optimization of a MEMS gyroscope for automotive applications, Mech. Syst. Sig.
Process. 128 (2019) 389–404, https://doi.org/10.1016/j.ymssp.2019.03.049.
[34] C. Surace, K. Worden, G. Tomlinson, Novelty Detection Approach to Diagnose Damage in A Cracked Beam, Proceedings of the International Modal
Analysis Conference – IMAC, 1, 947-953, 1997.
[35] E.L. Wilson, Three-Dimensional Static and Dynamic Analysis of Structures, Computers and Structures Inc, Berkely, 2002.
[36] K. Worden, Structural fault detection using a novelty measure, J. Sound Vib. 201 (1) (1997) 85–101.
[37] Z. Zembaty, P. Bobra, P.A. Bońkowski, S. Kokot, J. Kuś, Strain sensing of beams in flexural vibrations using rotation rate sensors, Sens. Actuators, A 269
(2018) 322–330, https://doi.org/10.1016/j.sna.2017.11.051.

17
L. Huras, Z. Zembaty, P.A. Bońkowski et al. Mechanical Systems and Signal Processing 151 (2021) 107396

[38] Z. Zembaty, S. Kokot, P. Bobra, Application of rotation rate sensors in an experiment of stiffness ‘reconstruction’, Smart Mater. Struct. 22 (2013) 077001.
[39] Z. Zembaty, M. Kowalski, S. Pospisil, Dynamic identification of a reinforced concrete frame in progressive states of damage, Eng. Struct. 28 (2006) 668–
681, https://doi.org/10.1016/j.engstruct.2005.09.025.
[40] W.-J. Yan, M.-Y. Zhao, Q. Sun, W.-X. Ren, Transmissibility-based system identification for structural health Monitoring: Fundamentals, approaches, and
applications, Mech. Syst. Sig. Process. 117 (2019) 453–482, https://doi.org/10.1016/j.ymssp.2018.06.053.

18

You might also like