Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Case Studies in Construction Materials 16 (2022) e01117

Contents lists available at ScienceDirect

Case Studies in Construction Materials


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

Operational modal analysis and finite element model updating of


ultra-high-performance concrete bridge based on ambient
vibration test
Siti Shahirah Saidin a, Sakhiah Abdul Kudus a, b, *, Adiza Jamadin a, b,
Muhamad Azhan Anuar c, Norliyati Mohd Amin a, Zainah Ibrahim d,
Atikah Bt Zakaria e, Kunitomo Sugiura f
a
School of Civil Engineering, College of Engineering, Universiti Teknologi MARA, Shah Alam 40450, Selangor, Malaysia
b
Institute for Infrastructure Engineering and Sustainable Management (IIESM), Universiti Teknologi MARA (UiTM), Shah Alam, Selangor, Malaysia
c
School of Mechanical Engineering, College of Engineering, Universiti Teknologi MARA, Shah Alam 40450, Selangor, Malaysia
d
Department of Civil Engineering, Faculty of Engineering, University of Malaya, 50603, Kuala Lumpur, Malaysia
e
Bridge Design Division, Public Work Department, 50582, Kuala Lumpur, Malaysia
f
Structures Management Engineering Laboratory, Department of Urban Management, Graduate School of Engineering, Kyoto University, Kyoto 615-
8540, Japan

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

Keywords: The advancement of technology for accurate and dependable monitoring and evaluation of cur­
Ambient vibration test rent bridge infrastructure conditions has become increasingly important in ensuring that bridge
Operational modal analysis structures operate safely. Operational modal analysis (OMA) has been used to solve various en­
Finite element analysis
gineering problems, including civil engineering structures, where OMA has been used to monitor
Modal identification
Modal updating
systems on a global scale. Structural health monitoring (SHM) is a method of observing and
Structural health monitoring assessing a process that involves sampling response measurements regularly to track changes in
the material and geometric properties of structures. However, knowledge of massive structural
health monitoring (SHM) data is poorly interpreted due to computational constraints and a lack of
data analysis methodologies. By determining the structure’s modal parameters, including the
natural frequency and mode shape, this study applied ambient vibration testing to acquire vi­
bration response measurements and thus determine the structure’s dynamic characteristics. A
comparison of three modal detection approaches (Frequency Domain Decomposition (FDD),
Enhanced Frequency Domain Decomposition (EFDD), and Stochastic Subspace Identification
(SSI)) was performed, and the mode shapes of each method were validated using the Modal
Assurance Criterion (MAC) value to verify the accuracy of the results. On the basis of experi­
mental data, a sensitivity-based updating method was used to justify and update the numerical
finite element (FE) model of bridge structures. The difference between the updated FE and that
measured for the first five dominant natural frequencies in this study was less than 5%. The
updated model’s minimum MAC value of 90% indicates that it was updated successfully. The
updated dynamic parameter of the first dominant natural frequency (3.348 Hz) was used to figure
out the structure’s serviceability vibration limit state in accordance with EN 1991–2. The result
was in the range that indicates that the bridge is safe for people to drive on.

* Corresponding author at: School of Civil Engineering, College of Engineering, Universiti Teknologi MARA, Shah Alam 40450, Selangor,
Malaysia.
E-mail address: sakhiah@uitm.edu.my (S.A. Kudus).

https://doi.org/10.1016/j.cscm.2022.e01117

Available online 2 May 2022


2214-5095/© 2022 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

1. Introduction

Bridging structures are vital components of the transportation network infrastructure. The bridge enables any obstacles to be
quickly overcome, bringing social communities together and stimulating the country’s economic growth. As industry and the popu­
lation grew in the mid-19th century, there was an increase in demand for the transportation network, which resulted in the con­
struction of more roads and bridges to cross rivers and valleys [1]. Every aspect of bridge safety and economics must be considered
when designing this vital transportation infrastructure component. Therefore, a diverse range of bridges is needed to be constructed.
Given that most bridges are made of reinforced concrete, the research focuses heavily on the new generation of ultra-high-performance
concrete (UHPC), which intends to provide an innovative solution to current bridge system problems [2–5].
A typical high-strength UHPC has been proposed premised on a composite concrete matrix consisting of fibre-reinforced cemen­
titious material constituted of Portland cement, silica fume (for reinforcement), mineral fillers, fine silica sand, superplasticiser, water,
and steel fibres [6–8]. Other supplementary cementitious materials used in the manufacture of UHPC, in addition to steel fibre,
encompass fly ash (FA), ground granulated blast furnace slag (GGBS), and rice husk ash (RHA). The high strength of up to 200 MPa
enables UHPC to be used as the primary material in structural bridge elements [9–12]. The distinctive feature of UHPC girders that can
span greater distances has met the design criteria of the bridge structure and the waterway in lowering the number of piers in the river
to reduce hydraulic defect structures. Furthermore, UHPC girders can be structured with smaller section sizes and no shear stirrups
[13–15]. As a result, construction time, labour, and consequently, costs were reduced. Among the world’s longest bridges, a footbridge
with a span of 120 m and a road bridge with a span of 100 m have been recorded [16,17]. Nevertheless, as UHPC bridge structures get
smaller and slenderer, they become more susceptible to external dynamic excitations like vehicle-induced vibrations. Consequently,
vehicle vibration loads play an essential role in developing long-span UHPC bridges, and the dynamic characteristics of these bridges
have been a prime concern for structural safety evaluation. Non-destructive testing (NDT) has been the most cost-effective and efficient
method of monitoring structural performance in response to these concerns. Additionally, this technique is extremely advantageous
since it causes no damage to existing bridge structures and does not disrupt traffic operations [18].
The research on ultra-high-performance composite structures is extensive, but it primarily focuses on the material’s mechanical
properties [19–21] and structural applications [22–25]. Additionally, significant research on the dynamic properties of UHPC
structures has been conducted [26–28]. However, the dynamic performance of UHPC bridges in the context of applications is still
sorely missing in the open literature [29–32]. This information is critical when using UHPC bridge structures, as they are more sus­
ceptible to vehicle-induced vibrations and are gaining popularity in the bridge industry. As a result, this paper presents ambient vi­
bration testing (AVT) performed on a single span of a 50 m length UHPC road bridge in Sungai Raia, Ipoh, Perak, Malaysia. Ambient
vibration interactions were measured at various locations on the bridge during normal traffic operations. The modal parameters of the
bridge were approximated using output-only modal detection approach like FDD, EFDD, and SSI [33]. Analytical finite element (FE)
modelling was used to determine the modal characteristics of the UHPC bridge. Uncertainties in FE simulations are common; thereby, a
refined and upgraded framework was designed to account for the differences between measured and assessed modal parameters.
This paper combined an ambient vibration-based approach, a FE model, and model updating to evaluate the structural effectiveness
of the UHPC road bridge. This method is more accurate and efficient than the traditional visual inspection method since the traditional
visual inspection strategy is highly subjective and does not offer a quantifiable measurement for vast and complex bridge structures.
Finally, the obtained modal parameters were checked and validated in accordance with the EN1991–2 standard [34] to certify the
efficiency limit of the UHPC bridge structure. This research adds to the novel development of vibration serviceability performance and
estimation of UHPC road bridges as additional dynamic effects (including resonance) criteria in EN1991–2 Clause 6.4.

2. Operational modal analysis

Operational modal analysis (OMA), also regarded as output-only modal analysis, is a method for determining a structure’s modal
properties by gathering vibration responses under unidentified input excitation scenarios. The ARTeMIS Modal Pro 7.0 version is one
of the commercially available OMA software packages. ARTeMIS has used several identification methods to validate the accuracy of
the results, along with Frequency Domain Decomposition (FDD), Frequency Domain Decomposition (EFDD), and automated Stochastic
Subspace Identification (SSI). Every one of these techniques is briefly explained in this section.

2.1. Frequency domain decomposition (FDD)

FDD is a widely known procedure for measuring the modal parameters of dynamic systems predicated on their response to external
excitation. It is a frequency domain technique that recognises higher-order modes with high precision [35]. The FDD technique, for
example, can assess closely coupled structural modes from a SHM system in the frequency domain. Non-parametric or parametric
frequency domain methods start by predicting an output spectrum or half-spectrum matrices based on measured dynamic responses.
The non-parametric FDD is a simple and straightforward technique for separating closely spaced modes. The modes are determined by
locating the spikes in singular value density (SVD) plots centred on the response’s power spectral density. No damping is calculated
because this FDD model focuses on a single frequency line from an FFT study, which limits the natural frequency estimate’s accuracy to
the FFT resolution. FDD was first evidenced by Brincker et al. [36]. The method’s principles, however, have already been applied to
the study of structures subjected to ambient excitation by Prevosto [37] and Correa and Costa [38], as well as the documentation of
modal parameters from FRF [39].

2
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 1. Sungai Raia, Kinta, Perak UHPC bridge. (a) Elevation view, (b) Front view (mm) [43], (c) Cross-section of UHPC U-grider [43].

2.2. Enhanced frequency domain decomposition (EFDD)

Unlike FDD, Enhanced Frequency Domain Decomposition (EFDD) is a parametric extension of FDD. EFDD provides a more accurate
estimate of natural frequencies, mode shapes, and damping. The Inverse Discrete Fourier Transform is used to invert the SDOF Power

3
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 2. Vibration measurement locations (m).

Fig. 3. Measurement set up. (a) 2 sensors attached at P1, (b) DAQ connected to the laptop set up.

Spectral Density function to the time domain after identifying a resonance peak (IDFT). The natural frequency is calculated as the
number of zero-crossings as a function of time, and the damping is calculated as the logarithmic decrement of the correlating SDOF
normalised autocorrelation equation. The SDOF function is estimated using the shape established by the prior FDD peak picking [40].

2.3. Stochastic subspace identification (SSI)

Time domain operations modal analysis techniques are based on the examination of response time histories or their correlation
functions. These techniques are usually best suited to dealing with noisy data and are the most effective time-domain identification
algorithms [41]. The SSI method’s proclivity for accurately classifying spaced modes, as well as their suitability for automation,
clarifies the reason of it getting more attention. Modal parameters are identified using stochastic processes, which means the structure
is excited by an unmeasurable input force and only output readings (e.g., accelerations) can be used to evaluate the structure’s modal
parameters. In these methods, the deterministic knowledge of the input is replaced by the assumption that the data is a realisation of a
stochastic process (white noise) [41]. The related modal parameters are identified using both output-only and data-driven SSI
methods. The raw time data is used in the data-driven SSI method, as is the correlation-based covariance-driven SSI method [42].
Unweighted Principal Components (UPC), Principal Components (PC), and Canonical Variate Analysis are the three algorithms used
for SSI (CVA). Nonetheless, the results of the PC and CVA are the ultimate focus of this study.

3. Ambient vibration test

The Sungai Raia bridge spans the Sg. Raia river on the federal route FT001 between Ipoh and Gopeng Road in northwestern
Malaysia. This is a new bridge built to broaden an original reinforced concrete bridge. The old bridge opened in 1962 and was
destroyed after 58 years of service. The monitored bridge is a 50-meter-long single-span ultra-high-performance concrete UHPC bridge
that is fully integrated and designed for SV 100 loading following BS EN 1991–2–2003. The bridge deck is 11.50 m wide, with one
traffic and motorcycle lane. The deck thickness is 22.50 cm, and the span is supported by three U-shaped precast post-tension girders
with a depth of 1.75 m and a sectional area of 15.24 mm2 embedded with 42 units of 270 ASTM A416 super strand wire, as presented in
Fig. 1.
Accelerometers with low frequency, high sensitivity, and low broadband noise levels are well suited for testing the ambient vi­
bration of large structures. The bridge was monitored using 15 single-axis accelerometers with a sensitivity range of 1000 mV/g
manufactured by PCB Piezotronics and installed at ten different locations along the bridge deck. The points were spaced 8.3 m apart,

4
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 4. Acceleration time series graph for point 1.

dividing the bridge deck into six cross-sections. Fig. 2 illustrates the locations of the ten points, namely P1 to P10. Each of the first five
points (P1-P5) is measured in two directions to represent vertical (+Z) and transverse (-Y) vibration responses. The remaining five
points (P6-P10) were kept from intercepting only vertical response (+Z) to capture the vertical bending mode.
The reliability of the measurement results was checked by calibrating each of the sensors. The accelerometers were linked to a data
acquisition (DAQ) unit and streamed to a laptop, allowing for capturing and displaying structure vibration signal response, as shown in
Fig. 3. The ambient testing was performed for 30 min at 100 Hz sampling rates to measure and record adequate vibration data for
modal parameter extraction.

4. Result and discussion

4.1. Modal identification

The responses from all 15 accelerometers mounted on the bridge’s deck were used to investigate the bridge’s vibration properties.
The modal parameters of the ambient vibration response or output-only where the unknown load input was defined using the FDD,
EFDD, and SSI modal identification algorithms. The graph in Fig. 4 indicates the time histories gathered in September 2020 for 30 min
at a sampling rate of 100 Hz and a time interval of 0.01 s. Each sensor channel’s measurement data capture 180,000 acceleration data
points. The vertical direction is slightly greater than the lateral direction, as shown in Fig. 4 (b). This is a common scenario for AVT on
bridges since the spectral density of modal forces in the vertical direction due to congestion is much higher than in the transverse
direction, primarily due to wind conditions. The vibration amplitudes in the transverse direction, 0.16 m/s2, were found to be 70% of
those in the vertical direction, 0.28 m/s2 for this UHPC bridge, and this was also observed in a previous study [44]. Natural excitations
such as human walking, wind loads, and particularly traffic loads (weight and speed of road vehicles) all significantly affect the vertical
vibration response. As evidenced by the testing, the peak acceleration appeared when the heavy vehicle crossed the bridge road at a
high rate of speed.
The experimental modal parameters were derived from the time histories collected during the experiment and processed using
commercially available post-processing software. The data were detrended to remove the mean value as well as any linear patterns in
the measurements that could tamper with signal processing and modal analysis. The 100 Hz sampling rate data was then decimated to
0–50 Hz to narrow the frequency range down to the frequency range of interest. These 30 min of data were segmented and processed
with a frequency resolution of 1024. Overlapping consecutive segments created a smooth singular value density (SVD) graph in the
frequency domain by 66%. The SVD method can decompose an original signal into a series of component signals linearly. It is a noise
reduction processing function that uses the study of singular value mutation features to determine several practical singular values for
complete reconstruction implementation signal noise reduction. The generated SVD graph was used to calculate the signal-to-noise
ratio (S/N) by comparing the favoured signal level to the background noise. According to Fig. 5, the three colour lines in the graph

5
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 5. Singular Value Decomposition (SVD) graph.

represent the most desired signal, which is essential for producing relevant information on modal data without unnecessary noise. The
figure strongly suggests that the noise floor is approximately 70 dB away from the DC. For this experimental test, the signal-to-noise
ratio (S/N) for the dominant peak is around 70 dB, indicating a good S/N for the OMA test. The minimum range for a suitable signal-to-
noise ratio in OMA is 30–40 dB [45].
The modal parameters were extracted from these measurements using ARTeMIS ProSoftware, a commercial post-processing
software. There are several operational modal analysis (OMA) algorithms, such as frequency domain decomposition (FDD),
enhanced frequency domain decomposition (EFDD), and stochastic subspace identification (SSI). The frequency-domain decompo­
sition (FDD) process extracts modal parameters from vibration measurements without the need for input load information [46].
Improved estimates can be obtained using the FDD method, which can identify closely coupled modes [47]. The EFDD and SSI methods
were used to validate and verify the FDD-derived modal frequencies and mode shapes. Fig. 6 shows a graph plot with some noticeable
peaks in the zoom in the frequency range 0–25 Hz, which correlate to various modes. As presented in Fig. 6 (a) and (b), two peaks at
10.5 Hz and 12.5 Hz were not selected as the preferred mode as they were noticed to be coupled and repeated modes. Integration of the
second vertical and first lateral bending modes caused the peak at 10.5 Hz. At the same time, there was a repetition mode of the first
vertical bending mode and a coupled mode with the first lateral bending mode at 12.5 Hz. This coupling mode was presumed to occur
when symmetric and antisymmetric pairs of mode shapes were combined [48]. This was caused by the placement of the sensors, with
the transverse sensor being positioned on one side of the bridge due to the symmetry of the bridge. The lateral mode shape generated in
the process of the coupling mode is inconsistent and complex.
Attributed to the reason that this study is concerned with capturing bending modes up to the third (3rd) order, the SVD graph
plotted contains only five peaks, each portraying five modes of interest with less complexity than the other five. The following fre­
quencies were chosen to represent only bending and torsion mode: 3.235 Hz, 6.548 Hz, 8.415 Hz, 13.336 Hz, and 16.852 Hz.
Meanwhile, the noise interruption stability diagrams and modes are illustrated in Fig. 6 (c) and (d) using the SSI-PC and SSI-CVA
algorithms. The SSI algorithm-generated nearly comparable findings in all modes of interest in FDD and EFDD, indicating that all
modes selected were stable and unnoisy modes.
From the operational modal analysis, the first five dominant modes and natural frequencies are summarised in Table 1. The first
modal frequencies acquired from the OMA are in the 3 Hz range, typical for a short span bridge with a 50 m span length, as determined
by previous 224 bridge testing. In general, bridge frequencies range from 2 to 4 Hz, except the long-span bridge’s frequency ranges
from 0 to 1 Hz [49]. This indicates the validity of the modes obtained because the variation in identified modal frequencies for
different algorithms is less than 1%.
The Modal Assurance Criteria (MAC) were also applied to compare the modal parameters acquired from all of the OMA methods
and to determine the relative accuracy of the various methods. The MAC is a metric that can be used to identify the association between
each obtained mode shape. The FDD method was selected as a point of comparison as it is well established among the most reliable and
user-friendly approaches for operational modal analysis. Each mode shape (FDD, EFDD, and SSI) is well-correlated and gives MAC
values close to 1, implying high correlation, as shown by the MAC graph in Fig. 7.

4.2. Temperature effect

Extensive studies have been carried out into the probability of temperature influencing the productivity and the structural
behaviour of the sensor. It is well known that environmental elements can have a significant impact on the effectiveness of a sensor’s
performance. As presented in Table 2, four measurements were taken over two days, with 30 min allotted for each data set. The
measured natural frequencies show a slight difference, with an average of 3.184 Hz on the first day (38.63 ◦ C) and 3.215 Hz on the
second day (38.1 ◦ C), indicating a 0.97% difference. The trend of natural frequency changes for the following 2nd mode and the rest
was quite similar. It demonstrates that the correlation between temperature and natural frequency is inversely proportional, with
higher temperatures lowering natural frequency values [50,51]. Nevertheless, as the temperature difference between different
measurements is in the range of 1–2%, it can be spotted that temperature changes slightly influenced the natural frequency value for
this study.

6
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 6. FDD, EFDD peak picking and SSI SVD graph. (a) FFD, (b) EFDD, (c) SSI-PC, (d) SSI-CVA.

7
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Table 1
Modal parameters based on various modal identification technique.
Mode Modal Identification technique

FDD (Hz) EFDD (Hz) SSI-PC (Hz) SSI-CVA (Hz)

1 1st vertical bending 3.235 3.235 3.268 3.273


2 1st torsion 6.548 6.548 6.505 6.535
3 2nd vertical bending 8.415 8.415 8.477 8.484
4 2nd torsion 13.336 13.336 13.285 13.301
5 3rd vertical bending 16.852 16.852 16.873 16.889

Fig. 7. MAC comparison graph for each method. (a) MAC correlation of FDD and EFDD, (b) MAC correlation of FDD and SSI-PC, (c) MAC cor­
relation of FDD and SSI-CVA.

Table 2
Natural frequencies from different measurement.
Date Measurement Time Average Temperature (◦ C) FDD (Hz)

30/09/2020 M1 11.40–12.10 40.2 3.125


M2 1400 − 1430 39.4 3.223
M3 1435–1505 38.6 3.174
M4 1510–1540 38.3 3.213
01/10/2020 M5 0930–1000 36.8 3.21
M6 1000–1030 39.8 3.208
M7 1030–1100 39.6 3.235
M8 1100–1130 40.2 3.208

8
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 8. Bridge finite element model.

Table 3
Mix design of UHPC [43].
Ingredient Mass (kg/m3)

UHPC premix 2100–2200


Superplasticizer 30 – 40
Steel Fiber 157
Free water 144
3% Moisture 30
Targeted W/B Ratio 0.15
Totak Air Void < 4%

Table 4
Material properties of superstructure bridge model [43].
Charactertistics Symbol Unit Standard Concrete UHPC

Specific Density ρ Kg/m3 BS EN 12390–7 2400 2500


Mean Cylinder compressive strength fcm MPa BS EN 12390–3 50 128
Mean cube compressive strength fcm,cube MPa BS EN 12390–3 65 145
Modulus of elasticity E GPa BS EN 12390–13 35 45
Poison Ratio V – BS EN 12390–13 0.2 0.2
Limit of elasticity under tension fctm,el MPa NF P18–470 4 >8

4.3. Numerical model of the tested structure

The dynamic analysis was carried out using a three-dimensional solid model of the Sungai Raia superstructure bridge, which
included the reinforced concrete deck and the UHPC girder. The finite element assessment was conducted using the commercial
modelling software ABAQUS 2017 version. Fig. 8 depicts a 3D model of a bridge superstructure. As stated in the as-built drawing issued
by the Ministry of Public Works Malaysia, Jabatan Kerja Raya and Voo et al. studies [43], the material features of the reinforced
concrete deck and UHPC were appointed in the finite element (Table 3, Table 4 and Table 5).

Table 5
Nominal stress and strain for strand 270 [52].
Nominal stress (N/m2) Nominal strain (N/m2)

1 0 0
2 1347000000 0.007
3 1496000000 0.008
4 1605000000 0.009
5 1675000000 0.01
6 1752000000 0.0125
7 1781000000 0.015
8 1799600000 0.0175
9 1815000000 0.02
10 1831000000 0.0225
11 1846000000 0.025
12 1862000000 0.0275
13 1862000000 0.03
14 1862000000 0.035

9
Table 6

S.S. Saidin et al.


Comparison of natural frequencies and appropriate mode shape of FEM with field testing.
MODE Finite Element Model (FEM) Operational modal analysis (OMA)

1st vertical
bending
(No. 1)

2nd
vertical
bending
(No. 3)
10

3rd vertical
Bending

Case Studies in Construction Materials 16 (2022) e01117


(No. 8)

(continued on next page)


Table 6 (continued )
MODE Finite Element Model (FEM) Operational modal analysis (OMA)

S.S. Saidin et al.


1st torsion
(No. 2)

2nd torsion
(No. 6)
11

1st lateral OMA did not yield any lateral modes. As per the outcomes, the lateral mode produced was a coupling mode that was
bending a combination of vertical and lateral modes, and the mode complexity mode in OMA is significant.
(No.5 No lateral mode was obtained from OMA. Based on the finding, the lateral mode produced was a coupling mode that
was a mix of vertical and lateral modes, and the mode complexity mode in OMA is high.

Case Studies in Construction Materials 16 (2022) e01117


2nd lateral
bending (No.
12)
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Table 7
Comparison of natural frequencies of FEM and OMA before model updating.
No. Mode Natural frequency (Hz) Difference, Δf (%) MAC (%)

fFEM fOMA

1 1st vertical bending 3.478 3.235 7.23 98.1


2 1st torsion 6.017 6.548 8.45 98.0
3 2nd vertical bending 9.201 8.415 8.92 95.3
4 2nd torsion 12.160 13.336 9.22 64.3
5 3rd vertical bending 17.135 16.852 1.67 86.8

Table 8
MAC diagonal before model updating.
FEM OMA

1 2 3 4 5

1 98.1 0.4 0.1 31.2 2.5


2 0.1 98.0 0.0 0.2 0.1
3 0.0 0.0 95.3 2.0 0.3
4 0.0 0.4 0.1 64.3 0.3
5 3.5 0.0 0.0 0.1 86.8

4.3.1. Reinforced concrete deck element


The element C3D8R, which is described as a three-dimensional, eight-node, second-order, reduced-integration element, was
applied to model the concrete deck members to significantly reduce the analysis’s computational power and run time. Truss elements
(T3D2), 3D and 2-noded, were used to model reinforcing rebars. Meanwhile, T25–125 bar reinforcement is utilised to reinforce the
bridge deck. The concrete members house the reinforced bar element. Dimensional and cross-sectional area modelling of all rein­
forcing rebars (longitudinal and transverse rebars) accompanied the reinforcement details mentioned by Voo et al. [43].

4.3.2. UHPC U-girder element


The U-girder was represented by 8-node solid linear hexahedral elements (C3D8R). The strand bar was created as T3D2 truss
elements and hosted in the UHPC U-girder solid element.

4.3.3. Constraint and interactions


The characteristic of embedded elements is defined as a collection of components involved within the host elements. This technique
is beneficial for simulating the bonding interaction of reinforcing rebars and the concrete that encircles them in reinforced concrete
structures. The fully bonded state is represented by limiting the translational degrees of freedom of the reinforcing rebars (embedded
elements) to those of the surrounding concrete (host elements). Using an embedded constraint, the 270-grade wire strands were
embedded in the UBG1750 girder using the solid-element concept. The longitudinally aligned nodes of the reinforced concrete deck
and UHPC girder were linked using a node to node tie constraint.

4.3.4. Boundary condition and loading


The fixed support at both ends of the bridge span is assigned for integral bridges, and rotation and translation are restricted at the
end span. The boundary condition was assigned to each node of the bridge deck support elements. Based on its physical and mechanical
properties, the purpose of the FE model in this analysis is to identify the dynamic properties of free/natural vibration, including natural
frequencies and corresponding mode shapes. As a result, no-load input was appointed to the FE model. The analysis type is modal
frequency analysis, and the mode extraction procedure is the Block Lanczos method.
The measured natural frequencies from field testing (fFEM OMA) were compared to the modal measurements derived from the FE
model natural frequency (fOMA FEM). Testing measurements provide useful information about the tested structure’s configuration. FE
models, on the other hand, can predict the dynamic behaviour of a structure under diverse loading, material type, size dimension, and
boundary conditions. Table 6 displays the FEM and OMA’s natural frequencies and mode shapes.
The result obtained from the fFEM and fOMA is then compared and evaluated the discrepancies. The discrepancies were quantified
by using simple relative error (Δf) of the FE model natural frequency (fFEM ) against the testing measured (fOMA ) as expressed in Eq. (1).
⃒ ⃒
⃒fOMA − fFEM ⃒⃒
Δf = ⃒⃒ ⃒x100 (1)
fOMA
According to the comparison results in Table 7, the percentage difference between FEM and OMA ranges from 3% to 13%. Given
that the maximum tolerable frequency discrepancy is 15% and the Modal Assurance Criterion (MAC) value is higher than 60%, the
outcomes of FEM before updating are considered reliable [53]. The MAC is a statistical indicator that is most prone to large distinctions
in mode shapes and relatively insensitive to minor differences. The correlated mode pairs between OMA and FEM were determined
using MAC values. As a result, the first and second lateral bending mode shapes were not subjected to FEM updating because the

12
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 9. Global Sensitivity matrix selected parameter respect to response. (a) Parameter 1, (b) Parameter 2, c) Parameter 3, d) Parameter 4.

minimum MAC requirement was not met and the mode from OMA was not stable. As the MAC outcome before updating is displayed in
Table 8, the FEM is updated for vertical bending and torsion.
When using finite-element modelling, the following factors may contribute to discrepancies between the bridge’s FEM and OMA: 1)
errors in the discretisation of the analytical model; 2) uncertainties in the geometry and boundary requirements; and 3) variations in
the material properties of the bridge [54]. Thus, a new finite element model is needed to accurately predict the structural health of the
bridge’s existing structures to conduct structural health monitoring.

4.4. Model updating

The bridge FE model must be updated to clarify any inaccuracies or uncertainties in input parameters like Young modulus, mass
density, poison’s ratio, and boundary conditions. A sensitivity-based updating method was used to resolve the inaccuracies. The
response and parameter selection drive the pre-process of sensitivity analysis and model updating. The target response was chosen
based on the natural frequencies and Modal Assurance Criterion (MAC) from field testing, with a 1% scatter for resonance frequencies
and a 10% scatter for MAC value. The precision of the evaluated resonance frequencies is relatively high than the accuracy of the
measured mode shape; thus, the scatter (%) frequencies are lower than the measured mode shape. The difference in frequencies
between mode pairs before model updating is less than 15%, and the MAC is 80% (Table 8). Therefore, it was necessary to ensure that
there was no more than a 5% difference between the updated model and the original model’s natural frequencies and that the MAC
value was at least 90%.
The parameters preferred should be reasonably sensitive and uncertain for the model to be effectively revised. A sensitivity analysis

13
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Table 9
Selected uncertain parameters for model updating.
Parameters Element Symbols Initial values Unit

1 U-Girder E 4.500E+ 10 N/m2


2 Concrete Deck E 3.500E+ 10 N/m2
3 U-Girder RHO 2.400E+ 3 Kg/m3
4 U-Girder NU 0.2 –

Table 10
Comparison of initial and updated values of selected parameters for model updating.
Parameters Element Symbols Initial values Updated value Difference, Δ (%)

1 U-Girder E 4.500E+ 10 4.209E+ 10 11.04


2 Concrete Deck E 3.500E+ 10 3.173E+ 10 9.35
3 U-Girder RHO 2.400E+ 3 2.130E+ 3 11.24
4 U-Girder NU 0.2 0.199 0.354

Table 11
Natural frequencies and Modal Assurance Criterion (MAC) of the UHPC bridge (Hz) after model updating.
Mode Natural frequency (Hz) Difference, Δf (%) MAC (%)

fFEM Updated fOMA

1st bending 3.348 3.235 3.51 98.7


1st torsion 6.281 6.548 4.16 98.5
2nd bending 8.561 8.415 1.72 97.5
2nd torsion 13.815 13.336 3.53 94.7
3rd bending 16.657 16.852 1.16 90.2

Table 12
MAC diagonal after model updating.
FEM OMA

1 2 3 4 5

1 98.7 0.3 0.0 0.2 2.6


2 0.1 98.5 0.1 0.1 0.1
3 0.0 0.0 97.5 0.0 0.2
4 0.0 0.1 0.0 94.7 0.3
5 2.6 0.0 0.1 0.0 90.2

is performed to aid in parameter selection for model tuning and eliminates ineffective (low sensitivity) parameters. Fig. 9 portrays a bar
chart of the global sensitivity matrix of the parameters chosen. The sensitivity matrix highlights the response sensitivity of selected
parameters (young’s modulus of the U-grider, young’s modulus of the concrete deck, mass density of the U-girder, and U-girder
poison’s ratio) (the first five natural frequencies and MAC value). Table 9 outlined the chosen parameters based on the properties of the
elements and materials.
The sensitivity analysis highlighted that the Young’s modulus of the UHPC U-girder, Fig. 9 (a), and the concrete deck, Fig. 9 (b),
were the variables most delicate to frequency modifications. For UHPC U-girder, the frequencies were less sensitive to material density
and poison’s ratio (Fig. 9 (c) and 9 (d)). Despite the low sensitivity of these parameters, they were still considered to be updating
parameters since it applies to the most critical structural components.
The global and local methods are used to select parameters. Instead of updating the global mass and stiffness independently for each
element, the international method can be easily constructed from the reference data, making it more efficient. The global parameter
was chosen to automatically update the parameters 1–4 of the uncertain (Table 10). The difference changes in the parameters are up to
11%, indicating that the material is uncertain in the initial model and showing good agreement with the sensitivity analysis result.
Modifications in material parameters lead to response parameters such as natural frequencies and MAC alterations. After the model
was updated, the FE model’s chosen parameter was upgraded to improve the FE model’s accuracy, permitting the deviation between
the resonance frequencies to be minimised and the MAC to be maximised. Table 11 compares the MAC values of experimental and
globally updated FE frequencies. The first five dominant natural frequency differences between updated and evaluated models were
less than 5%, and the updated model’s least MAC value of 90% indicates that the model was successfully updated [33]. Table 12 and
Fig. 10 reveal that the MAC diagonal shape is acceptable, and the MAC correlation met the acceptable minimum range. Moreover,
Fig. 11 shows a strong correlation between the updated FE and the experimental natural frequencies, with a regression value of
R= 0.9885. The automated model updating and the global model updating were in excellent accordance, suggesting that the

14
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Fig. 10. Modal Assurance Criterion (MAC) correlation of UHPC bridge before and after model updating. (a) Mac before model updating, (b) MAC
after model updating.

Fig. 11. Correlation of updated FE and experimental frequency.

automated model updating was appropriately implemented.


In addition, Table 13 also indicates the relationship between the finite element and experimental mode shapes after the updated
process is completed. The red line represents the experimental mode shape for five cross-section areas in the figure, and the blue 3D
model represents the updated finite element model. The five corresponding selected modes show that the FEM and OMA results
correlate well, implying that the model has been successfully updated.

4.5. Natural frequency based on theoretical

The theoretical calculation is then used to validate the modal parameter, such as natural frequency, derived from FEM and
experimental OMA. Dahleh [55] describes natural frequency theory as follows in Eq. (2):
√̅̅̅̅̅̅
EI
ωn = (βn l)2 (2)
ρl4

Natural frequencies are interpreted by Eq. 2 in contexts of the modulus of elasticity (E), the moment of inertia (I), the mass density
(M), and the span length (L) [34]. The natural frequency equation is for lateral vibration regarding the typical end condition exhibited
in Table 14.
For the integral bridge case, the clamped-clamped condition (βn l2 = 22.4, 61.7 and 121) was applied to determine the theoretical

15
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Table 13
Mode shape comparison between FEM and test after model updating.
MODE FEM-field test mode shape after model updating

1st bending

1st torsion

2nd bending

2nd torsion

3rd bending

Table 14
(βn l)2 based on boundary condition [56].
Beam configuration (βn l)2 Fundamental (βn l)2 Second mode (βn l)2 Third mode

Simply supported 9.87 39.5 88.9


Cantilever 3.52 22.0 61.7
Free-free 22.4 61.7 121.0
Clamped-clamped 22.4 61.7 121.0
Clamped-hinged 15.4 50.0 104.0
Hinged-free 0 15.4 50

natural frequency value, ωn . Then, the ωn in rad/sec is converted to 1/2π Hz to compare with the FEM and experimental finding. The
correlated data of experimental and FEM based on theoretical were computed in Fig. 12 and show acceptable where the regression
value, R2, is close to 1.0, and the updated model was validated with the theoretical value where the percentage difference is between
1% and 5%.

16
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

20
R² = 0.9883

Natural Frequency, f (Hz)


15

10

0
1 2 3
Lateral bending mode
theoretical OMA FEM

Fig. 12. Natural frequency comparison between theoretical, OMA and FEM.

Fig. 13. Limits of bridge natural frequency (Hz) as a function L∅ for integral bridge type, EN1991–2:2003 [41].

The natural frequencies are proportional to the structure’s stiffness, EI, and span length. Fundamentally, the natural frequency is
evaluated by the stiffness of the structure, and the correlation is directly proportional; as the structure’s stiffness decreases, so does the
natural frequency. It can be concluded that as the structure endures some degree of damage or deterioration, a decrease in natural
frequency may lead to alterations in stiffness [56]. Thus, the modal parameter like natural frequency has proven to be an excellent
indicator for assessing the current structural condition of bridges.

4.6. EN1991-2:2003 natural frequency standard

The first natural frequency determined from the updated model is validated in accordance with EN1991–2:2003 [57]. The girder
span length of 50 m is considered the L ∅ for an integral bridge with simply supported girders and slabs (along with steel beams
embedded in concrete) (Table 6 in EN1991–2:2003). Fig. 13 reveals that the 3.348 Hz frequency falls within the acceptable range and
is in compliance with the standard. The natural frequency of the structure is also inversely proportional to the span length, as displayed
in the illustration. The longer the design span, the lesser the value of the natural frequencies. Lower natural frequency values can be
found in the 0–1 Hz range, particularly for long bridge structures. Hence, the natural frequency of UHPC bridges of 60, 70, 80, 90, and
100 m was predicted using the updated model and verified with the European standard (Table 15). This finding demonstrates that the

17
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

Table 15
Updated FEM frequency for various span length with respect to the EN1991–2:2003 frequency max and min limitations.
Span length FEM frequency (Hz) Theoretical frequency (Hz) Percentage difference (%) Classification

10 m 10.167 10.012 1.63 Lies in the range


50 m 3.348 3.241 3.25 Lies in the range
60 m 2.573 2.487 3.39 Lies in the range
70 m 2.127 2.083 2.09 Lies in the range
80 m 1.476 1.364 7.88 Does not lies in the range
90 m 1.168 1.078 8.01 Does not lies in the range
100 m 0.933 0.892 4.49 Does not lies in the range

span ratio is another important parameter that influences vibration frequencies. For long single-span bridges, UHPC offers advantages
because of its unique mechanical behaviour and high durability. The DURA project, for example, consisted of a 100-meter-long precast
segmental box girder of the UHPC at Batu 6, Gerik, Perak [58]. Nonetheless, this study found that the serviceability of the UBG1750
UHPC U-girder design is only capable of serving up to 70 m of UHPC single-span bridge. Since the U-Girder has an 80 m span length
and above, further research is needed to expand the UHPC material’s capabilities on a single-span bridge.

5. Conclusion

The dynamic characteristics of the Sg. Raia UHPC bridge (natural frequencies and mode shape) were determined through the use of
finite element modelling and ambient vibration testing. Based on the observations above, the following conclusions are reached:

i. The FE model and ambient vibration testing of the bridge identified the first five appropriate bending and torsion modes with
natural frequencies ranging from 3 to 17 Hz, leading to development of an updated model.
ii. The successful FE model update shows that it is possible to effectively update a large bridge infrastructure model by modifying
the inaccuracies or uncertainties in the input parameters based on output-only modal identification results.
iii. The updated dynamic parameter of the first dominant natural frequency (3.348 Hz), which falls within the acceptable natural
frequency range according to EN1991-2:2003, was indicated to be in agreement with the design.
iv. Modal parameters like structural stiffness and natural frequency can be applied to assess current structural bridge condition,
with changes in natural frequency being correlated with modifications in structural stiffness and indicating damage or
deterioration.

Therefore, the method developed in this study has potential application in assessing the structural integrity of real-world bridges on
a larger scale, as well as in setting industry standards in the future. A non-destructive method of monitoring existing bridges without
interfering with traffic is provided, making it a viable, cost-effective, and time-consuming option for the authority or structural owner.
It is probable to develop a predictive model of the serviceability of a structure by using more data from different periods. This can be
accomplished by performing additional experimental testing regularly over at least three to four distinct periods.

Ethical approval

This article does not contain any studies with human participants or animals performed by any of the authors.

Declaration of Competing Interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing in­
terests: Sakhiah Abdul Kudus reports financial support was provided by Universiti Teknologi MARA Malaysia.

Acknowledgement

The authors would like to express their gratitude to Universiti Teknologi MARA, Shah Alam, Malaysia for supporting this publi­
cation via research grant 600-RMC 5/3/GRR (004/2020).

References

[1] A. Gonzalez, M. Schorr, B. Valdez, A. Mungaray, Bridges: structures and materials, Ancient and Modern. Infrastructure Management and Construction,
Intechopen, 2020.
[2] R. Zhao, Y. Yuan, X. Wei, R. Shen, K. Zheng, Y. Qian, C. Yu, Review of annual progress of bridge engineering in 2019, Adv. Bridge Eng. 1 (1) (2020) 1–57.
[3] Tayeh, B.A., Aadi, A.S., Hilal, N.N., Bakar, B.H. A., Al-Tayeb, M.M., Mansour, W.N. (2019). Properties of ultra-high-performance fiber-reinforced concrete
(UHPFRC)—a review paper. International Symposium On Green And Sustainable Technology (ISGST2019).
[4] N.K. Baharuddin, F. Mohamed Nazri, B.H. Abu Bakar, S. Beddu, B. A. Tayeh, Potential use of ultra high-performance fibre-reinforced concrete as a repair
material for fire-damaged concrete in terms of bond strength, Int. J. Integr. Eng. 12 (9) (2020) 87–95.
[5] B.A. Tayeh, B.H.A. Bakar, M.A.M. Johari, Y.L. Voo, Utilization of ultra-high performance fibre concrete (UHPFC) for rehabilitation – a review, Procedia Eng. 54
(2013) 525–538.

18
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

[6] O. Gunes, S. Yesilmen, B. Gunes, F.J. Ulm, Use of UHPC in bridge structures: material modeling and design, Adv. Mater. Sci. Eng. 2012 (2012).
[7] M. Abdul-Rahman, A.A. Al-Attar, H.M. Hamada, B. Tayeh, Microstructure and structural analysis of polypropylene fibre reinforced reactive powder concrete
beams exposed to elevated temperature, J. Build. Eng. 29 (2020), 101167.
[8] A.-A. Alyaa A, B. Mazin A., M. Hussein H., A. Bassam T., Investigating the behaviour of hybrid fibre-reinforced reactive powder concrete beams after exposure to
elevated temperatures, J. Mater. Res. Technol. (2019).
[9] M. Amin, A.M. Zeyad, B.A. Tayeh, I.S. Agwa, Effect of ferrosilicon and silica fume on mechanical, durability, and microstructure characteristics of Ultra high-
performance concrete, Constr. Build. Mater. (2022).
[10] A.S. Faried, S.A. Mostafa, B.A. Tayeh, T.A. Tawfik, Mechanical and durability properties of ultra-high performance concrete incorporated with various nano
waste materials under different curing conditions, J. Build. Eng. 43 (2021), 102569.
[11] M. Amin, A.M. Zeyad, B.A. Tayeh, I. Saad Agwa, Engineering properties of self-cured normal and high strength concrete produced using polyethylene glycol and
porous ceramic waste as coarse aggregate, Constr. Build. Mater. 299 (2021), 124243.
[12] M. Amin, B.A. Tayeh, S.A. Ibrahim, Effect of using mineral admixtures and ceramic wastes as coarse aggregates on properties of ultrahigh-performance concrete,
J. Clean. Prod. (2020), 123073.
[13] W. Mansour, M. Sakr, A. Seleemah, B.A. Tayeh, T. Khalifa, Development of shear capacity equations for RC beams strengthened with UHPFRC, Comput. Concr.
27 (5) (2021) 473–487.
[14] W. Mansour, B.A. Tayeh, Shear behaviour of RC beams strengthened by various ultrahigh performance fibre-reinforced concrete systems, Adv. Civ. Eng. 2020
(2020).
[15] W. Mansour, M. Sakr, A. Seleemah, B.A. Tayeh, T. Khalifa, Bond behavior between concrete and prefabricated Ultra High-Performance Fiber-Reinforced
Concrete (UHPFC) plates, Struct. Eng. Mech. 81 (3) (2022) 305–316.
[16] N.M. Azmee, N. Shafiq, Ultra-high performance concrete: from fundamental to applications, Case Stud. Constr. Mater. 9 (2018), e00197.
[17] Y.L. Voo, S.J. Foster, C.C. Voo, Ultrahigh-performance concrete segmental bridge technology: toward sustainable bridge construction, J. Bridge Eng. 20 (8)
(2015), B5014001.
[18] S.T. Lin, Y. Lu, M.M. Alamdari, N.L. Khoa, Field test investigations for condition monitoring of a concrete culvert bridge using vibration responses, Struct.
Control Health Monit. 27 (10) (2020), e2614.
[19] E. Fehling, M. Schmidt, J. Walraven, T. Leutbecher, S. Frönlich, Ultra-high performance concrete UHPC: fundamentals, Des., Ex., Beton-Kal. (2013).
[20] M. Shafieifar, M. Farzad, A. Azizinamini, Experimental and numerical study on mechanical properties of Ultra High Performance Concrete (UHPC), Constr.
Build. Mater. 156 (2017) 402–411.
[21] J. Li, Z. Wu, C. Shi, Q. Yuan, Z. Zhang, Durability of ultra-high performance concrete–a review, Constr. Build. Mater. 255 (2020), 119296.
[22] Y. Zhu, Y. Zhang, H.H. Hussein, G. Chen, Flexural strengthening of reinforced concrete beams or slabs using ultra-high performance concrete (UHPC): A state of
the art review, Eng. Struct. 205 (2020), 110035.
[23] H.G. Russell, B.A. Graybeal, H.G. Russell, Ultra-high performance concrete: a state-of-the-art report for the bridge community (No. FHWA-HRT-13-060). United
States, Fed. Highw. Adm. Off. Infrastruct. Res. Dev. (2013).
[24] M. Zhou, W. Lu, J. Song, G.C. Lee, Application of ultra-high performance concrete in bridge engineering, Constr. Build. Mater. 186 (2018) 1256–1267.
[25] J. Xue, B. Briseghella, F. Huang, C. Nuti, H. Tabatabai, B. Chen, Review of ultra-high performance concrete and its application in bridge engineering, Constr.
Build. Mater. 260 (2020), 119844.
[26] P.R. Prem, M. Verma, A.R. Murthy, P.S. Ambily, Smart monitoring of strengthened beams made of ultrahigh performance concrete using integrated and
nonintegrated acoustic emission approach, Struct. Control Health Monit. (2021), e2704.
[27] X. Zhang, X. Li, R. Liu, C. Hao, Z. Cao, Dynamic properties of a steel–UHPC composite deck with large U-ribs: experimental measurement and numerical
analysis, Eng. Struct. 213 (2020), 110569.
[28] M. Elsayed, B.A. Tayeh, M. Abou Elmaaty, Y. Aldahshoory, Behaviour of RC columns strengthened with Ultra-High Performance Fiber Reinforced concrete
(UHPFRC) under eccentric loading, J. Build. Eng. 47 (2022), 103857.
[29] Chin, W.J., Kim, Y.J., Cho, J.R., & Park, J.S. (2012). Dynamic characteristics evaluation of innovative UHPC pedestrian cable stayed bridge.
[30] Tej, P., Kněž, P., Vráblík, L., & Kolísko, J. (2017). Modal analysis of cable-stayed UHPC bridge. In MATEC Web of Conferences (Vol. 107, p. 00007). EDP
Sciences.
[31] L. Deng, S. Zou, W. Wang, X. Kong, Fatigue performance evaluation for composite OSD using UHPC under dynamic vehicle loading, Eng. Struct. 232 (2021),
111831.
[32] J. Wang, J. Liu, Z. Wang, T. Liu, J. Liu, J. Zhang, Cost-effective UHPC for accelerated bridge construction: material properties, structural elements, and structural
applications, J. Bridge Eng. 26 (2) (2021), 04020117.
[33] R. Brincker, C. Ventura, Introduction to operational modal analysis, John Wiley & Sons, 2015.
[34] EN 1991–2 Eurocode 1: Actions on structures - Part 2: Traffic loads on bridges.
[35] C. Rainieri, G. Fabbrocino, E. Cosenza, Some remarks on experimental estimation of damping for seismic design of civil constructions, Shock Vib. 17 (4–5)
(2010) 383–395.
[36] Brincker, R., Zhang, L., Andersen, P. (2000) Modal identification from ambient responses using frequency domain decomposition, in Proceedings of the IMAC
18, International Modal Analysis Conference, San Antonio, USA.
[37] Close, M., Prevosto (1982). Algorithmes D′ Identification des Caractéristiques Vibratoires de Structures Mecaniques Complexes, Ph.D. Thesis, Université de
Rennes I, France.
[38] M.R. Correa, A.C. Costa, Dynamic tests of the bridge over the Arade River, in: J.A. Fernandes, L.O. Santos (Eds.), Cable-stayed Bridges of Guadiana and Arade,
Book, LNEC, 1992.
[39] C.Y. Shih, Y.G. Tsuei, R.J. Allemang, D.L. Brown, Complex mode indicator function and its application to spatial domain parameter estimation, Mech. Syst.
Signal Process. 2 (4) (1988) 367–377.
[40] Jacobsen, N.J., Andersen, P., & Brincker, R. (2006). Using enhanced frequency domain decomposition as a robust technique to harmonic excitation in
operational modal analysis. In Proceedings of ISMA2006: international conference on noise & vibration engineering. Katholieke Universiteit.
[41] W.X. Ren, Z.H. Zong, Output-only modal parameter identification of civil engineering structures, Struct. Eng. Mech. 17 (3–4) (2004) 429–444.
[42] M.A. Anuar, A.A. Mat Isa, Z. AR, Critical experimental issues of cracked aluminum beam in operational modal analysis, J. Mech. Eng. JMechE 6 (2018) 211–225.
[43] V.Y. Lei, B. Nematollahi, A.B.M. Said, B.A. Gopal, T.S. Yee, Application of ultra high performance fiber reinforced concrete–the Malaysia perspective, Int. J.
Sustain. Constr. Eng. Technol. 3 (1) (2012) 26–44.
[44] M.Q. Feng, Y. Fukuda, Y. Chen, S. Soyoz, S. Lee, Long-term structural performance monitoring of bridges, Tech. Rep. Calif. Dep. Transp., 2006-UCI-02 (2006).
[45] R. Brincker, C. Ventura, Introduction to Operational Modal. Analysis, first ed., Wiley, New York, 2015.
[46] R. Brinker, L. Zhang, P. Andersen, Modal identification of output-only system using frequency domain decomposition, Smart Mater. Struct. 10 (3) (2001)
441–455.
[47] Otte, D., Ponseele, P.V. D., Leuridan, J. (1990), Operational shapes estimation as a function of dynamic loads, in Proceedings of 8th International Modal Analysis
Conference, Society for Experimental Mechanics, Orlando, FL, 413–21.
[48] M. Hey Leung, J. Corcoran, A probabilistic method for structural integrity assurance based on damage detection structural health monitoring data, Struct.
Health Monit. (2021), 147592172110388, https://doi.org/10.1177/14759217211038881.
[49] Idris, N., Boon, K., Kamarudin, A. and Sooria, S. (2016). Ambient Vibration Test on Reinforced Concrete Bridges. MATEC Web of Conferences, 47, p.02012.
[50] G.-D. Zhou, T.-H. Yi, A summary review of correlations between temperatures and vibration properties of long-span bridges, Math. Probl. Eng. 2014 (2014)
1–19.
[51] Y. Cai, K. Zhang, Z. Ye, C. Liu, K. Lu, L. Wang, Influence of temperature on the natural vibration characteristics of simply supported reinforced concrete beam,
Sensors 21 (12) (2021) 4242.

19
S.S. Saidin et al. Case Studies in Construction Materials 16 (2022) e01117

[52] R.K. Devalapura, M.K. Tadros, Stress-strain modeling of 270 ksi low-relaxation prestressing strands, PCI J. 37 (2) (1992) 100–106.
[53] J.M.W. Brownjohn, P.Q. Xia, H. Hao, Y. Xia, Civil structure condition assessment by FE model updating: methodology and case studies, Finite Elem. Anal. Des.
Vol. 37 (No. 10) (2001) 761–775.
[54] M. Huang, W. Guo, H. Zhu, L. Li, Dynamic test and finite element model updating of bridge structures based on ambient vibration, Front. Archit. Civ. Eng. China
2 (2008) 139–144.
[55] W.T.T. Dahleh, M. D, Theory of Vibrations with Applications, fifth ed., Pearson, USA, 2014, pp. 271–273. England.
[56] A. Jamadin, Z. Ibrahim, M.Z. Jumaat, E.S. Ab Wahab, Effect of high-cyclic loads on dynamic response of reinforced concrete slabs, KSCE J. Civ. Eng. 23 (3)
(2019) 1293–12931301.
[57] BS EN1991–2:2003 Eurocode 1: Actions on structures - Part 2: Traffic loads on Bridges, section 6.4.4, page 77.
[58] Y. Voo, S. Foster, C. Voo, Ultra-high-performance concrete segmental bridge technology: toward sustainable bridge construction, J. Bridge Eng. 20 (2015) 8.

20

You might also like