Armijo 2024 Integration of Railway Bridge Struc
Armijo 2024 Integration of Railway Bridge Struc
Article
Integration of Railway Bridge Structural Health Monitoring into
the Internet of Things with a Digital Twin: A Case Study
Alberto Armijo * and Diego Zamora-Sánchez
TECNALIA, Basque Research and Technology Alliance (BRTA), Astondo Bidea, Edificio 700,
48160 Derio, Spain; [email protected]
* Correspondence: [email protected]
Abstract: Structural health monitoring (SHM) is critical for ensuring the safety of infrastructure such
as bridges. This article presents a digital twin solution for the SHM of railway bridges using low-cost
wireless accelerometers and machine learning (ML). The system architecture combines on-premises
edge computing and cloud analytics to enable efficient real-time monitoring and complete storage
of relevant time-history datasets. After train crossings, the accelerometers stream raw vibration
data, which are processed in the frequency domain and analyzed using machine learning to detect
anomalies that indicate potential structural issues. The digital twin approach is demonstrated on an in-
service railway bridge for which vibration data were collected over two years under normal operating
conditions. By learning allowable ranges for vibration patterns, the digital twin model identifies
abnormal spectral peaks that indicate potential changes in structural integrity. The long-term pilot
proves that this affordable SHM system can provide automated and real-time warnings of bridge
damage and also supports the use of in-house-designed sensors with lower cost and edge computing
capabilities such as those used in the demonstration. The successful on-premises–cloud hybrid
implementation provides a cost effective and scalable model for expanding monitoring to thousands
of railway bridges, democratizing SHM to improve safety by avoiding catastrophic failures.
Keywords: structural health monitoring (SHM); railway bridges; digital twin (DT); machine learning
(ML); MLOps; low-cost MEMS accelerometers; vibration-based monitoring; wireless sensor networks
(WSNs); hybrid computing; building information modeling (BIM)
Citation: Armijo, A.; Zamora-
Sánchez, D. Integration of Railway
Bridge Structural Health Monitoring
into the Internet of Things with a
1. Introduction
Digital Twin: A Case Study. Sensors
2024, 24, 2115. https://doi.org/ Railway bridges are indispensable components of global transportation networks
10.3390/s24072115 and play a crucial role in facilitating the seamless movement of goods and passengers.
Ensuring their structural integrity is not just a matter of maintenance but a critical safety
Academic Editor: Jiawei Xiang
imperative. Traditional methods of inspecting this vital infrastructure through visual
Received: 24 January 2024 surveys are not only resource-intensive but also sporadic and disruptive to daily operations.
Revised: 8 March 2024 Recognizing these challenges, this article introduces an innovative approach to structural
Accepted: 22 March 2024 health monitoring (SHM) of railway bridges, integrating the principles of the Internet
Published: 26 March 2024 of Things (IoT) with a digital twin framework. This solution employs low-cost wireless
sensors and leverages advanced machine learning algorithms to automate the monitoring
process, offering real-time insights into the structural health of railway bridges.
Section 1 provides insightful background on railway bridges and highlights the im-
Copyright: © 2024 by the authors.
portance of SHM for these critical assets. It also summarizes previous research efforts that
Licensee MDPI, Basel, Switzerland.
independently explored components like low-cost sensing, energy harvesting, digital twins,
This article is an open access article
distributed under the terms and
and machine learning for vibration-based SHM of bridges. Finally, the section introduces
conditions of the Creative Commons
the proposed digital twin solution for SHM of railway bridges.
Attribution (CC BY) license (https:// Section 2 outlines the key objectives and performance metrics of our study, including
creativecommons.org/licenses/by/ achieving high classification accuracy for detecting abnormal vibration patterns using
4.0/).
1.1. Brief Background of Railway Bridges and Importance of Structural Health Monitoring
The transport system—in particular, the railway system—is essential to our daily lives,
and plays a fundamental role in the social and economic development of a region and
a country [1]. The transport networks in the Basque Country (Spain) and in Europe are
highly developed but face a growing problem of aging as, for example, 35% of European
railroad bridges are more than 100 years old [2]. Also, many of these bridges, especially the
relatively old ones, are being pushed beyond their designed physical capacities due to the
increased speed, axle load, length, and travel frequency of new trains and transportation
needs [3–6].
Climate change introduces uncertainties, since previously unseen climatic conditions
and phenomena tend to occur more frequently and unexpectedly, making these structures
particularly vulnerable [7]. In addition, bridges, given their relatively low level of structural
redundancy, are generally at risk of collapse in the event of significant damage caused by
deterioration, resulting in serious economic losses, interruptions in the normal course of
people’s lives, and, in the worst case, irreparable tragedies.
Recently, there have been several major bridge collapses around the world, including
the Morandi Bridge failure in Genoa in 2018 [8] and the Nanfang’ao Bridge collapse in Yilan
County in Taiwan in 2019, among others [9]. These incidents show the potential dangers
when critical structural damage goes undetected. As a result, frequent and proactive
monitoring of bridge structural health is essential to avoid such catastrophic failures. Minor
issues like cracking or loose connections can gradually worsen over time if not spotted early.
Taking a broad definition, structural health monitoring (SHM) is the process of assess-
ing the state of health of a structure based on data from instrumentation installed on the
structure [10,11]. It is a process that can be divided into different phases depending on the
criterion: some do it in three, such as in [12], although we distinguish four, since manage-
Sensors 2024, 24, 2115 3 of 30
aspects of our proposal. In this sense, the novel technique presented in the aforementioned
article offers high accuracy with low-range sensors and damage detection sensitivity down
to 0.01 Hz frequency shifts, whereas in our approach, we rely on clustering that addresses
the variability in identification due to multiple factors. This technique could also be suitable
to be implemented in the digital twin platform presented in this research through cloud
processing of the obtained records.
The article [22] proposes an automated framework to classify anomalies (i.e., drift,
distortion, outlier, anomaly, bias, etc.) in the time domain and assess the current state of the
structure, while in our approach, we work in the frequency domain, proposing clustering
for anomaly detection after the identification of frequencies in free vibration, which also
allows us to identify and discard faulty measurements. If a permanent discard were to
occur, a no-data alert would also be issued, and we would have the corresponding signals
in the time domain to perform a more detailed study or even accommodate algorithms
such as the one proposed in the aforementioned paper since our digital twin system or
architecture is flexible enough to include other algorithms and is capable of receiving and
storing data in the time domain with high sampling frequencies.
The work in [23] offers a new approach to damage identification based on the ex-
traction of continuous time series of autoregressive (AR) coefficients from deformation
measurements on a railway bridge, but it is based on fiber optic technology, i.e., expensive
instrumentation, while the application of the present work is based on low-cost accelerom-
etry after the passage of a train. In any case, the core of our work is a digital twin system
with a middleware component that can be readily adapted to other types of sensors with
any physical magnitude provided that communication can be implemented at least from a
PC connected in situ, while high-level algorithms, such as the one proposed in the work
cited above, could be included in our processing layer.
The article [24] offers a bibliographic review of how energy harvesting technologies
can provide sustainable power sources for wireless sensor networks (WSNs) deployed
on bridges. The SHM systems implemented on bridges are mostly based on WSN. Solar,
thermal, wind, and vibrational energy harvesting are all examined as ways to overcome the
limitations of battery-operated sensor platforms. However, [24] was confined to an energy
perspective and did not investigate how the data from such sensors could be utilized for
automated SHM powered by simulations and analytics.
The work in [25] discusses that MEMS sensors are miniature in size and have lower
cost and weight than conventional wired alternatives. These advantages make MEMS
sensors better suited for permanent installation over many years of continuous infrastruc-
ture monitoring. The article also provides a bibliographic review regarding commonly
used machine learning techniques, both classical and deep learning methods, for bridge
structural health analysis using sensor data. But it identified high computational costs and
model performance as limitations to practical cloud-based implementations for large-scale
infrastructure monitoring. Edge computing is suggested as a potential solution, but it is
not implemented.
The article [26] provides a review of machine learning algorithms that have been
successfully applied in SHM, specifically in the domains of vision-based and vibration-
based SHM. In this regard, this paper leverages a vibration-based approach. However,
other, more advanced AI powered deep learning algorithms could be implemented using
our demonstrated workflow.
In addition, our approach offers the added benefit of building information modeling
(BIM) [27] contextualization of sensors and measurements within an architecture adaptable
to other algorithms and sensors. Indeed, for this study, we designed and developed our
own low-cost devices with edge computing capabilities compatible with the digital twin
system. However, we prefer to present this digital twin as a system compliant with any
sensor and any machine learning algorithm as well as adaptable to all types of structural
configurations thanks to the geometric BIM contextualization of the bridge geometry and
the instrumentation. Secondly, it implements sustainable solar energy harvesting in the full
SHM solution, building on the potential shown in [24]. And thirdly, it delivers a hybrid
edge–cloud machine learning pipeline to make large-scale analytics financially feasible,
addressing the barriers called out in [25,26].
Specifically, our digital twin integrates inexpensive IoT acceleration sensors (equip-
ment costs around EUR 300) with MQTT [28] connectivity, on-premises fog computing,
cloud big data and machine learning services, and a visualization application. Compared
to [21], the sensor data are augmented by their contextual placement within a digital twin
model of the bridge for enhanced structural insights, with particular potential for mon-
itoring local variables (such as strain) through the use of other types of sensors beyond
accelerometers. This article applies the digital twin concept conforming to the definition
given in [20] and comprises the next features:
• Simulation: SHM + IoT + BIM;
• Learning: AI;
• Management: DS (provides decision support).
The article [29] presents a pre-trained network with synthetic data, i.e., supervised
with finite element models, which, based on deep learning techniques, could provide
a fast response for damage identification if integrated with real-time monitoring, and
which would be computationally suitable to be included in the digital twin system we
present, thus extending its simulation performance for model inclusion (adding the MOD
and DIAG function in the simulation capabilities according to the structural digital twin
conceptualization presented in [20]).
Moreover, powering the wireless sensors using solar energy harvesting realizes a
self-contained system, enabling the sustainable sensor networks envisioned by [24]. And
by leveraging both real-time on-premises edge processing and cloud machine learning, our
solution overcomes the prior constraints around computational costs described in [25,26],
demonstrating affordable analytics scaling.
To handle the high-throughput vibration data generated by the accelerometers, the
digital components of the structural health monitoring system were deployed in a hybrid on-
premises and cloud architecture [30]. For real-time data ingestion and analysis, on-premises
middleware was implemented. This on-premises system ingests the 500 Hz. sampled data
streamed from the accelerometers on the bridges via MQTT and runs real-time machine
learning algorithms to detect anomalies in the vibration patterns. To supplement this
real-time analysis, the historical vibration data are also regularly forwarded to a cloud
platform for longer-term storage and batch analysis. Storing and processing the entire
high-frequency vibration dataset solely in the cloud would be prohibitively expensive due
to large data volumes. By leveraging an on-premises system for time-critical analytics
combined with cloud storage and batch processing, the railway operator can cost-effectively
monitor the health of its bridges in real-time while also building up a knowledge base of
historical structural dynamics data.
• Achieve ≥ 95% classification accuracy for detecting abnormal vibration patterns that
indicate potential structural issues using only low-cost wireless sensors;
• Enable scalable large-scale monitoring across thousands of bridges through a cloud
and on-premises edge computing architecture costing 80% less than traditional wired
sensor networks;
• Reduce installation and maintenance complexity by over 50% compared to typical
SHM systems by utilizing self-contained wireless sensors with battery/solar power.
Performance in these areas will quantify and demonstrate the improvements of the dig-
ital twin approach over current SHM practices in rail infrastructure monitoring in terms of
precision, affordability, practicality, and democratization. The quantification is carried out
through showcasing a fast, reliable, and cost-effective remote damage detection system for
bridge structures that integrates the measured data and the alerts generated for the bridge
into the Internet of Things (IoT). The system aims to move from a reactive to a proactive
approach in bridge maintenance, replacing basic inspections with an automated process,
to take a first step towards the smartization of this infrastructure and its management as
Industry 4.0 [31] assets in a generalized way.
A pilot case of the system was implemented in a real environment for a railway bridge
in the Basque Country. For this purpose, following a vibration signal data type approach
and the application of a wireless sensor network (WSN) platform [24], an optimized sensor
plan was developed for the structure, and the measurement information was remotely
processed, enhancing its usability through its synthesized visualization on dashboards
accessible from a geometric model in the cloud.
One major contribution of our present work is the deployment and evaluation of the
digital twin capabilities in an operational context on a real railway bridge, which also
proves the viability of AI-powered digital twins using low-cost wireless IoT sensors. For
structural health monitoring of infrastructure like this bridge, the digital twin is powered by
machine learning algorithms instead of traditional physics-based simulations. Data-driven
approaches, e.g., [32], are widely used in SHM in both the time and frequency domains,
but they are not usually based on low-cost IoT sensors leveraging the clustering-based
approach of this work. This approach is particularly well aligned to work in real-time
using Eigenfrequencies with uncertainty in their identification (not only environmental and
operational but also due to sensor and measurement limitations). We first train the digital
twin, as is the common practice in unsupervised data-based approaches, on vibration
data collected from the bridge under known normal conditions. This allows the machine
learning model to learn the patterns of vibration that correspond to normal structural
dynamics. The trained digital twin model is then connected to real-time vibration data
streamed from accelerometers on the actual bridge. By analyzing these vibrations using
its trained machine learning algorithms, the digital twin can detect anomalies that deviate
from the learned ”normal” patterns. These anomalies may indicate potential structural
problems not discoverable through visual inspection.
Most of the traditional methods of operational modal analysis, or modal identification
without input measurement, work with high-sensitivity and high-price accelerometers,
such as the force-balance type, but in this case, we work with low-cost MEMS sensors, as
the only valid measurable output is the one produced by the passage of the train, so that
the free vibration after the exit of the train from the structure is a signal that contains only
the natural frequencies of the bridge.
The digital twin model can be re-trained over time as more sensor data are collected
to improve its accuracy. The machine learning approach provides a data-driven way
to monitor bridge health without relying on complex physics simulations. By detecting
vibration anomalies, the digital twin can provide early warning of damage so that repairs
can be made before catastrophic failure occurs.
This work establishes a replicable and cost-effective methodology for real-time railway
bridge monitoring that can be extended to large infrastructure networks. The approach
demonstrates reliable high-frequency data collection using MQTT communication be-
Sensors 2024, 24, 2115 7 of 30
tween low-cost sensors and cloud platforms. Compatibility with commercial off-the-shelf
acquisition modules enables flexible adoption with existing monitoring hardware.
A hybrid on-premise and cloud architecture processes the high-volume sensor streams
using open-source tools for edge analytics and cloud machine learning. Mosquitto, Node-
RED, and time-series databases handle real-time needs, while cloud services provide
scalable data storage, batch processing, and model management. The architecture is sensor-
agnostic and adaptable to new data sources.
Additionally, a realistic digital twin integrates the real-time sensor data with bridge geom-
etry models and AI-generated health insights for enhanced situational awareness. Interactive
dashboards connect the physical infrastructure state with digital monitoring outputs.
Overall, this pilot study proves the real-world viability of transitioning from costly
manual inspections to continuously automated AI-powered infrastructure health moni-
toring. By demonstrating a pragmatic digital twin system architecture using affordable
off-the-shelf components, this work enables scalable structural monitoring to improve
railway operations, maintenance planning, and passenger safety.
Figure 1. Sensor and communication layer, including a local WiFi network generated by the gateway
and an MQTT publishing broker.
cluster assignment is returned to the middleware and stored. Any anomalies or changes in
the typical cluster distribution can indicate a potential structural issue requiring further
inspection. The on-premises database, in addition to the real-time inference results, registers
contextual sensor data such as temperature.
The edge layer acquires high-frequency sensor data, the on-premises middleware (A)
handles real-time processing needs, while the cloud (B) provides big data and machine
learning capabilities. The digital twin application (C) fuses sensor data with the bridge
information model to bring monitoring insight to users.
Sensors 2024, 24, 2115 10 of 30
In the data engineering workspace, Databricks was used to process the raw FFT data
from the accelerometers. Spark clusters running on Databricks consolidated and cleaned
these data, ultimately clustering the FFT spectra into three labeled groups representing
different vibration patterns. For data transformation, the medallion architecture [51] was
used. This transformed the raw data into curated silver and gold datasets to make them
suitable for training machine learning models.
The machine learning component of Databricks leveraged the AutoML [52] feature
to automatically train and evaluate different algorithms on the labeled FFT data. The
best-performing algorithm was selected and registered into the MLflow [40] tracking server.
MLflow enabled MLOps on Databricks by managing the model lifecycle, including model
versioning, staging, and production deployment. Trained models were first tested in a
staged environment and then promoted to production deployment on a serverless and
scalable Azure Kubernetes service. The production models were exposed as real-time
inference endpoints that could be called by the on-premises middleware system.
Figure 3. Railway bridge used as pilot at Torrelarragoiti industrial area, exploited by ETS Euskal
Trenbide Sarea/Basque Railway Network (Zamudio, Basque Country, Spain).
The 17 m single-span bridge chosen as the pilot for this structural health monitoring
system represents a typical medium-length span used for urban bridges. In this kind
of single-span bridges, deterioration issues eventually emerge from accumulated traffic,
material aging, and exposure to climatic conditions.
Sensors 2024, 24, 2115 12 of 30
In order to record as clearly and accurately as possible the first three vibration frequen-
cies, an accelerometer was placed close to the span center (to contrast the first longitudinal
bending mode and, as a local variable, the acceleration of the point of maximum displace-
ment, since this acceleration can also be controlled by a direct threshold), an accelerometer
was placed at one-fourth of the span (between the span center and support, approximately
where the maximum amplitude of the third modal form of vibration is located, correspond-
ing to a second longitudinal bending, to which the accelerometer at the span center is
theoretically blind), and an accelerometer was placed on the other beam to characterize the
torsion (second theoretical mode, which might not always be excited by the passing of the
centered load). However, this last added sensor ended up being removed before the end of
the installation, as the concrete slab shielded the coverage provided by the WiFi antenna
on the other side of the beam. That is, only one of the main girders was instrumented,
which was sufficient for the demonstration or functionality test of the system parts once
assembled, which is the objective of the pilot test. The installation was completed in about
3 h, but it may be carried out faster on more accessible bridges. It was necessary to make a
cut in the road under the bridge and use a flatbed truck to access the flange of the main
girders. It was decided to place the gateway and the accumulators on the flange and to
place the sensors next to the web. According to the usual practice of instrumentation
of structures, this is a suitable position for the measurement of accelerations and main
vibration frequencies.
It is important to emphasize that it is necessary to record measurements from bridges
for further analysis (a complete dynamic and/or more detailed study can be done by an
expert after downloading data from the platform) and to have a power outlet available.
Failing that, it is necessary to install solar panels and their respective accumulators, since
the measurement equipment generally cannot withstand many hours without power when
high measurement and transmission rates are required. IoT equipment that has autonomy
for long periods of time using only batteries usually measures and transmits data every
several minutes, while for the pilot experience, up to 500 data points per second were being
recorded for train passages (achieved by implementing communication through the MQTT
protocol and a 4G network).
Regarding data quality, the span of the bridge resulted in being a bit short, and the
free vibration of the bridge was restricted to less than 8 s. This meant that the part of the
signal (after the train exit) corresponding to the free vibration had to be precisely captured.
To solve this, and since the train units always cross at the same speed, a manual time
adjustment of the capture was configured in the middleware to ensure that a significant
free vibration signal was captured for each crossing.
Another issue came from the differences in temperatures from one day to another. The
sensors were sensitive to the temperature and summed an offset in the signal, generating
false triggering due to detection of signals that exceeded the RMS thresholds and activating
the capture of invalid signals. In addition, under certain climate circumstances, some noise
was induced due to interference that occasioned the same triggering issue. The solution
comprised regular remote calibration of the MQTT sensors when notable temperature
changes were forecast.
Concerning the selection of the on-premises and cloud infrastructures, a key challenge
was to find the right balance for leveraging both platforms cost-effectively. While cloud
services excel at data storage, batch processing, and machine learning, they can become
prohibitively expensive. For this bridge monitoring use case, the raw sensor data rates
were too high to process solely in the cloud. By handling real-time ingestion, parsing,
and FFT analysis on-premises, the data forwarded to the cloud were reduced by orders of
magnitude. This avoided excessive cloud data ingress charges. However, the cloud was still
leveraged for its strengths, such as cheap storage, distributed batch data processing, and
on-demand machine learning model training and deployment. These tasks are challenging
to implement on-premises due to the maintenance costs and configuration efforts of these
components. To scale the solution to many bridges, the on-premises components allow
Sensors 2024, 24, 2115 13 of 30
cheap real-time analytics, and the cloud then enables aggregated analytics by consolidating
data from all bridge deployments. Careful testing was conducted to find the optimal split
between real-time processing and batch cloud analytics. This balance minimized costs
while still providing a robust, low-latency structural health monitoring system.
4. Experimental Methodology
4.1. Generic Methodology
This sub-chapter introduces a comprehensive, generic methodology for structural
health monitoring (SHM) of railway bridges that leverages low-cost IoT sensors, hybrid
edge–cloud processing, finite element model (FEM) calibration, unsupervised learning to
identify normal vibration patterns, generation of synthetic damage data, and supervised
model training and validation. The goal is to enable accurate, cost-effective, and scalable
SHM that can be replicated for other pilot bridges beyond the specific one presented in this
manuscript. The proposed ideal framework consists of seven key steps:
1. Characterization of the railway bridge;
2. Sensor deployment and data acquisition;
3. Data preprocessing;
4. Unsupervised learning;
5. Model calibration and synthetic data generation;
6. Supervised learning;
7. Validation and continuous learning.
While not all seven steps were fully implemented for the pilot bridge in this work, the
overall methodology provides a complete reference that other researchers can adapt and
build upon to achieve robust SHM using affordable sensing and advanced analytics. The
specific application and customization of this generic approach for the pilot railway bridge
is detailed subsequently in Section 4.2.
By presenting this methodology, the aim is to establish a replicable foundation for data-
driven SHM that fuses IoT, AI, and digital twins to deliver a proactive, automated solution
for assessing bridge health at a fraction of the cost of traditional inspection-based or wired
monitoring techniques. This will allow infrastructure managers to scale real-time damage
detection across their entire bridge networks: enabling timely maintenance interventions
that prevent catastrophic failures and extend structure lifespans.
train loading. The sensors should be ruggedized with IP67 rating for outdoor conditions
and integrated with solar panels and rechargeable batteries for self-powered operation.
In our work, we implemented an optimized sensor installation plan by placing the
accelerometers at the best locations on the bridge based on the FE analysis. The sensors
connect to a local 4G gateway on the bridge through a WiFi mesh network, enabling
redundant communication paths. The gateway streams the high-frequency vibration data
to an on-premises middleware system using the MQTT protocol over 4G/5G networks.
Triggers were set to activate data capture only during train crossing events in order to
conserve bandwidth and battery life.
ings. To cope with this fact, a clustering algorithm was used to separate the measurements
into a reduced number of groups or cases. Basically, depending on the passing of the train,
some frequencies are not identifiable after the departure of the vehicle because they have
not been sufficiently excited, or signals with higher noise and/or insufficient duration of
free vibration are generated for a correct processing. These latter signals are discarded,
since the identification of frequencies with accelerations of free vibration is more unstable
with these more limited sensitivity sensors (lower than class A ones). That is, a series of
clusters can be separated to generate three different clusters. With clustering, different
regions of normality can be easily defined for the training of anomaly detection so that
once the algorithm has been trained, at the passage of each new train, the membership of
each of the clusters can be evaluated.
Figure 4 shows a time-series plot of free vibration after a passing train. The signal is fur-
ther processed with an FFT algorithm for frequency domain transformation and classification.
Figure 4. Time-series plot of the free vibration after the crossing of a train.
Figure 5 shows a plot of the free vibration FFT of the previous figure. Clustering is
performed based on the patterns of these modal frequencies for the characterization of the
normal situation of the bridge.
Figure 5. Frequency plot of the free vibration after the crossing of a train.
To validate this characterization, the free vibrations were processed after the passing
of a train to obtain a vector with the natural frequencies of the structure of the pilot
research case. These frequencies are approximately in agreement with the frequencies of
the Ansys Mechanical APDL [53] finite element model, knowing that this is only indicative
because there are no drawings of the structure and was no access permission to carry out
exhaustive measurements. Thus, it was verified with regard to the frequency domain that
Sensors 2024, 24, 2115 17 of 30
the measurements provided by the equipment are valid and that the bridge does not differ
appreciably from its theoretical behavior, i.e., as it is physically.
After assigning some parameters that could not be measured directly within acceptable
values (Young’s modulus of steel = 210,000 MPa, Young’s modulus of concrete = 35,000 MPa,
density of steel = 7850 Kg/m3 , density of concrete = 2400 Kg/m3 , and web thickness of the
non-standard steel girder = 65% of the measured flange), the values obtained by the finite
element model for the main three vibration modes are as follows:
• 5.71 Hz for the first mode (bending);
• 15.11 Hz for the second mode (torsion);
• 20.84 Hz for the third mode (second order bending).
The corresponding modal shapes are shown in Figure 6 (Ansys Mechanical APDL).
Figure 6. Modal shapes of the Ansys Mechanical APDL for the pilot bridge. The boundary conditions
are those of a simply supported beam (applying the restraints at the bottom flange end of the steel
girders in line with how the structure is actually supported).
4.3. How a Digital Twin Detects Structural Changes from Sensor Data
Over the 2 years of testing, the system collected around 4300 valid samples of bridge
vibration spectra during train crossing events. Table 1 shows a characterization of the
collected FFT data.
Sensors 2024, 24, 2115 18 of 30
Using k-means clustering on this FFT data, the digital twin consistently identified
three distinct clusters representing normal vibration patterns. Figure 7 illustrates the
distribution of the three clusters in a 3D space, along with a descriptive analysis of the data.
The relative sizes of the three clusters did not change as more data were accumulated over
time, confirming that the system learned the three intrinsic vibration patterns characterized
by the clusters.
Tables 2 and 3 show the distribution of the cases into clusters, the description of the
cluster center positions, and the statistical characterization of the three clusters.
Figure 8 summarizes the distribution of the total samples across these three clusters
identified under normal operating conditions, providing insight into the prevalence of
different vibration signatures in the raw acceleration data.
Sensors 2024, 24, 2115 19 of 30
Std.
Cluster Mean Minimum Maximum
Deviation
principal_1 3 5.75 0.18 4.64 6.84
2 5.67 0.11 5.62 6.35
1 5.69 0.11 5.62 5.86
principal_2 3 13.93 11.17 0 102.83
2 14.38 10.64 3.17 74.95
1 97.86 6.86 73.49 100.83
principal_3 3 7.86 8.11 0 35.4
2 71.18 23.11 50.29 119.87
1 23.6 19.99 0 86.67
According to the provided data, the 4323 cases were distributed as follows:
• Cluster 1 contained the fewest cases at 190 cases.
• Cluster 2 had 659 cases.
• Cluster 3 was the largest group with 3474 cases.
The dominance of Cluster 3 indicates that nearly 80% of the observed vibration patterns
fell into this category. Cluster 2 had the next highest portion at 15% of the cases, while
Cluster 1 represented about 5% of the measured spectra. This distribution shows that most
of the vibration data (95%) were classified into Clusters 2 and 3, which likely represent
the fundamental natural frequency of the bridge and low-order resonances. The fewer
members of Cluster 1 suggest it corresponds to less common transient or higher-order
vibration phenomena.
The clusters represent distinct vibration frequency patterns observed during train
crossing events. Clusters 3 and 2, with a peak frequency around 5 Hz, likely correspond
to the fundamental bending mode of the bridge span, and the peak frequency of 14 Hz
corresponds to the torsion mode. Cluster 1 probably captures higher-order resonances such
as high-order bending modes, including transient vibrations without clear modal patterns.
Sensors 2024, 24, 2115 20 of 30
By separating the raw acceleration signals into these characteristic clusters, the digital twin
model establishes baseline behavior profiles for the bridge.
The relative size of the clusters quantifies the prevalence of different vibration sig-
natures identified by the machine learning model under normal operating conditions.
Tracking changes in the cluster distributions over time then enables assessing shifts in the
structural response that may require further inspection. This cluster analysis demonstrates
the digital twin can characterize the normal response in terms of Eigenfrequencies from the
raw acceleration signals.
Under the assumption used in other works [32], it is possible to generate damage
in the computational model and consider the variation ratio it causes in the vibrational
parameters. Calculating the ratio—the relative change between the initial (healthy) and the
damaged scenario—eliminates the effect of modeling error and uncertainty for validation.
That is, the calculated variation ratio is subsequently applied to the real signals, assuming
that it is constant over the short period of time considered for testing. With this approach,
we built realistic damaged data that include the variability of experimentally measured
signals and use them to validate the performance of the anomaly detection algorithm.
Thus, given that there is no precise information and that the deterioration of the bridge
could be associated more with its general aging, a synthetic damage scenario was generated
with a 10% loss in the modulus of elasticity of the girder’s steel, thus obtaining an average
variation ratio (decrease) of the Eigenfrequencies of around 3%. This is not intended to
be a recreation of a real damage scenario but, rather, provides an order of magnitude of
variations to test the method.
As a consequence, new vibration modes were calculated from this simulated FEM state
by applying the variation ratio of 3% between the damaged and normal vibration modes to
the empirical sensor data clusters to create three corresponding synthetic damaged data
clusters. The normal and synthetic damaged labeled datasets were combined into one
dataset with a total of 8646 rows of data associated with the six clusters: three normal
operation clusters and three corresponding damaged data clusters. The clusters were la-
beled as: ok_cluster_1, ok_cluster_2, ok_cluster_3, damaged_cluster_1, damaged_cluster_2,
and damaged_cluster_3. This consolidated 8646 row dataset was split into 60/20/20
train/test/validation partitions.
Leveraging MLflow capabilities, the cloud machine learning model was trained using
AutoML to classify any new vibration data into one of these six clusters. Databricks
AutoML was leveraged to automatically evaluate different machine learning algorithms
and hyperparameters. The classification models tested included decision trees, random
forests, logistic regression, XGBoost, and LightGBM. An MLflow experiment was launched
that ran various combinations of these algorithms with tuned hyperparameters. At the
conclusion of the experiment, the best-performing model was selected based on accuracy
metrics. The hyperparameters of the LightGBM model selected were:
• colsample_bytree = 0.6317331055500884;
• lambda_l1 = 0.1459367945385852;
• lambda_l2 = 0.30720685012169846;
• learning_rate = 0.03691431372131188;
• max_bin = 264;
• max_depth = 5;
• min_child_samples = 36;
• n_estimators = 551;
• num_leaves = 7;
• path_smooth = 95.01320854755053;
• subsample = 0.5638297006431046;
• random_state = 190,645,121.
The optimal model proved to be a LightGBM classifier automatically trained using the
scikit-learn Python library. LightGBM can rapidly train classification models, and with its
blend of performance, accuracy and scalability, it has become a popular machine learning
Sensors 2024, 24, 2115 21 of 30
package suitable for a wide range of predictive modeling applications from analytics
competitions to production systems [54]. The structural health monitoring system utilizes
this classifier due to its high performance in training models on large vibration datasets to
accurately classify normal versus abnormal bridge conditions. The evaluation metrics for
the chosen LightGBM model were:
• Training data and splitting:
– The model was trained on 8646 rows of data;
– A 60/20/20 train/validation/test split was used;
– This means 5188 rows for training and 1729 rows each for validation and test sets.
• Best iteration (281) and stopped iteration (286):
– The model achieved its best validation performance after 281 boosting rounds;
– Training was stopped early at iteration 286, which was close to the best iteration;
– This indicates the early stopping callback worked well to prevent overfitting.
• Test metrics:
– Test log loss: 0.0120—very low, indicates high accuracy on unseen test data;
– Test ROC-AUC: 0.9999—near perfect discrimination ability on the test set.
• Training metrics:
– Training log loss: 0.0099—extremely low, model fits training data remarkably well;
– Training ROC-AUC: 0.9999—near perfect discrimination on training data;
– Potential for some overfitting, but this is not concerning given test performance.
• Validation metrics:
– Validation log loss: 0.0171—low, but higher than training, as expected;
– Validation ROC-AUC: 0.9999—excellent discrimination on validation data;
– Valid_0 log loss: 0.0171—same as overall validation for the first fold.
In summary, the model achieves very high accuracy, with near perfect discrimination
as measured by the ROC-AUC on both the validation and held-out test sets. The early
stopping callback effectively prevented overfitting. The low log loss values indicate the
model is well calibrated for its predicted probabilities. The consistency between validation
and test metrics is a good sign that the model generalizes well to new data.
By leveraging AutoML to automatically find the best classification algorithm through
hyperparameter tuning, we were able to build an accurate machine learning model for
detecting anomalies in the vibration data using the labeled training clusters. The model
serves as the predictive engine behind the digital twin’s structural health analysis capabil-
ities. By continuously analyzing the stream of new accelerometer data, the model could
detect when the vibration patterns deviated from the normal clusters, i.e., shifting from
ok_cluster_1 to damaged_cluster_1 with a probability >99.9%, accounting for a detection
resolution capable of detecting a 10% synthetic reduction in the steel’s elastic modulus.
To quantify the extent of the deviation, we defined a “damage index”, which is easily
interpretable by the end user, as the probability that a vibration data sample does not belong
to any of the three known normal clusters and belongs to the corresponding damaged
cluster. The higher this index, the more dissimilar the vibrations were from patterns seen
during normal operation.
Figure 9 shows a dashboard implementation with the “damage index” or probability
of failure after train passing. The reliability of this result increases with the number of data
points collected and the number of variables included.
These clusters established a profile of expected normal structural dynamics. By
continuously analyzing new data, any vibrations that deviated from the learned clusters
could indicate a change in the bridge’s condition.
For example, changes in vibration magnitudes (for a fixed input, i.e., the train) detected
in certain frequency bands could suggest a loss of structural stiffness, as would a reduction
in the main Eigenfrequency (abscissae of the FFT peaks) values. Also, new vibration peaks
Sensors 2024, 24, 2115 22 of 30
appearing could imply damage to the structure, like the prolongation of the vibration after
a train crossing (due to changes in structural damping).
While controlled damage testing was not possible on the active railway bridge, the
long-term monitoring data itself revealed seasonal shifts in vibration patterns due to factors
like temperature. The digital twin was able to adapt and maintain high accuracy despite
these operational variations. To cope with these seasonal shifts, the middleware and cloud
platform implemented a mechanism that filtered out the outliers in the vibration registries
that fell well outside the expected ranges. The correction of this issue was carried out
manually by accessing the sensors’ remote interfaces and adjusting the sensors’ offset.
Regarding user interaction and visualization, a BIM model of the bridge, generated by
RDT Ingenieros, and the sensors was built. The model enables a common data environment
wherein the geometric, structural, and sensor data are accessible and linked together as
a digital twin. Included in the features of the visualization system is a train crossing
visualization, as shown in Figures 10 and 11, for which the free vibration of a crossing is
shown in real-time.
Figure 10. Interactive digital twin of the bridge from a BIM model (via IFC file) including real and
historical sensor data on top of user-selected elements.
Another set of functionalities includes a system for sending alerts based on thresholds
configured by expert users and a dashboard with the probability of failure after a train
passing. In essence, the digital twin learned an allowable range of normal vibrations over
long-term monitoring through machine learning. This enables it to detect abnormal patterns
Sensors 2024, 24, 2115 23 of 30
that are potentially linked to structural changes. The system provides a data-driven method
for continuously assessing bridge health without requiring baseline physics models.
Figure 11. Interactive digital twin of the bridge from a BIM model (via IFC file) including an index to
access the properties of user-selected elements.
processed data (10% of the size of the raw data) would cost only $690 per month, resulting
in a 90% reduction in cloud storage costs.
The digital twin’s machine learning algorithms, which are trained on historical vibra-
tion data, achieved a classification accuracy of ≥99.9% for detecting the abnormal vibration
patterns indicative of structural issues. The LightGBM model, optimized using AutoML
techniques, demonstrated excellent performance in distinguishing between normal and
anomalous vibration signatures. The model’s high accuracy enables early detection of
potential damage before it becomes visually apparent during inspection.
Compared to state-of-the-art methods, the proposed digital twin solution offers several
key advantages:
1. Cost-effectiveness: achieves lower costs than traditional wired sensor networks
through the use of low-cost wireless sensors and edge–cloud computing.
2. Scalability: enables large-scale monitoring across thousands of bridges with minimal
additional infrastructure costs.
3. Ease of deployment: reduces installation and maintenance complexity through the
use of self-contained, battery/solar-powered wireless sensors.
4. High accuracy: attains outstanding classification accuracy for detecting abnormal
vibration patterns using rapid advanced machine learning techniques.
5. Actionable insights: provides continuous, automated structural health assessment
and generates maintenance recommendations to prevent failures.
IoT applications compared to using only cloud or only edge computing. The key enablers
are local data processing at the edge, intelligent workload distribution, dynamic resource
allocation, and cost-aware optimization algorithms. The evaluation results show that the
proposed edge–cloud approach achieves significant cost savings compared to cloud-only
or edge-only deployments.
The findings from the referenced articles support the target objectives and statements
of our article. The high classification accuracy achieved by the enhanced SVM algorithms,
particularly SVM-EN with 87.2% accuracy for simple beams, demonstrates the potential
of machine learning techniques for detecting abnormal vibration patterns using low-cost
wireless sensors. The cost and complexity reductions associated with wireless sensor net-
works and the benefits of a hybrid edge–cloud architecture further validate the feasibility of
enabling scalable large-scale monitoring across thousands of bridges while reducing instal-
lation and maintenance complexity. By quantifying these improvements and comparing our
results to the existing literature and systems, we can effectively demonstrate the advance-
ments and benefits of our proposed digital twin approach for SHM of railway bridges.
4. Ability to continuously learn and adapt over time as more real-world data accumu-
lates: The validated AutoML machine learning model serves as an initial baseline,
but it does not remain static. By storing all historical vibration data from the bridge
sensors in a cloud data lake, the system gains access to an ever-growing dataset.
At regular intervals, this expanding dataset is leveraged to retrain and update the
machine learning model, steadily enhancing its accuracy through additional training
on real examples from the field. This continuous learning capability allows the model
to automatically adapt to gradual evolutions in the bridge’s structural dynamics, such
as effects from seasonality, aging of materials, and other factors. Thus, the digital twin
does not remain frozen but instead dynamically evolves its understanding of normal
versus abnormal vibration patterns as conditions change over the bridge’s lifetime.
5. Visualization centered on a realistic digital twin with location and access to the
information of each sensor that is represented as the physical reality and has assigned
IDs: That is, IoT is integrated as a digital twin connecting streaming data, AI generated
alerts, and the physical geometry of the bridge.
The hybrid architecture allows real-time edge analytics combined with cloud machine
learning over historical data for an efficient, scalable structural health monitoring system.
The middleware handles real-time needs, while the cloud provides flexible big data and
machine learning capabilities. This architecture proved highly scalable, as additional
bridges can be onboarded by deploying standalone on-premise systems for each. The
centralized cloud platform aggregates and analyzes data from all bridges as needed.
The hybrid on-premises–cloud architecture was key to overcoming the implementation
challenge of applying cost effective and replicable machine-learning-based monitoring to
high-frequency sensor data from the infrastructure.
The successful pilot proves the field viability and paves the way for expanded produc-
tion deployments. Railway operators can obtain a reliable, low-cost solution to continuously
monitor their bridges and detect issues early to avoid disruptions.
This pilot demonstrates two significant benefits of the digital twin approach for
structural health monitoring. First, it enables automated, continuous assessment of bridge
integrity using only affordable wireless sensors. Second, the on-premises–cloud architecture
provides a scalable implementation model that can be expanded to monitor thousands
of bridges.
6. Conclusions
This pilot study demonstrated the capabilities of a digital twin solution for real-time
monitoring of railway bridge structural integrity using low-cost accelerometers. The digital
twin model was driven by vibration data streamed from the sensors to detect anomalies
through machine learning algorithms to indicate potential damage. The hybrid edge–
cloud architecture enabled an efficient and scalable implementation. The edge middleware
handled high-speed data ingestion, preprocessing, and real-time FFT analysis, while the
cloud provided storage, batch processing, and machine learning capabilities.
The digital twin’s machine learning algorithms were able to potentially detect and
localize damage from changes in structural vibration patterns before visual inspection. This
structural health monitoring approach based on a digitally twinned machine learning model
provides an automated way to continuously assess bridge safety. It generates actionable
insights on maintenance needs to prevent infrastructure failures and to avoid disasters.
For future work, we aim to extend this successful railway bridge demonstration to
broader production deployments. The low cost and ease of implementation make the digi-
tal twin concept scalable to enable monitoring of all critical bridges across a transportation
network. By detecting issues early, this system has the potential to dramatically improve
infrastructure safety and prevent catastrophic failures through affordable, large-scale struc-
tural health evaluation.
While the digital twin and machine learning approach shows promise for monitoring
other types of infrastructure as well, we plan to first improve the fidelity of the digital
Sensors 2024, 24, 2115 27 of 30
twin model using advanced physics simulations and expanded sensor data for the railway
bridge use case. We will also research ways to better quantify the extent, localization, and
progression of any damage detected.
One promising infrastructure application is in wind towers. These massive structures
undergo important stresses and lack extensive instrumentation in the structure itself. A
digital twin could simulate wind tower structural dynamics to pinpoint damage using
limited sensor data. In this regard, the authors are exploring partnerships to develop digital
twin monitoring systems for wind towers: both onshore and offshore. By detecting issues
early, costly shutdowns and repairs could be avoided while ensuring these renewable
energy assets operate safely.
While the current system architecture utilizes middleware for preprocessing and cloud
machine learning, future work will shift more of the analytics onto intelligent sensors
that provide energy harvesting and edge computing characteristics. The authors are
developing custom sensors that can process FFT and vibration analysis directly on the
sensor hardware using embedded machine learning chips. These sensors utilize e-peas [61]
MPPT algorithms, state-of-the-art solar panels from Voltaic Systems [62], and Li-ion super
capacitors for energy harvesting and storage. Edge Impulse will be used to process the data
at the edge: progressively moving the middleware functionality from the fog to the edge.
The sensors are based on ESP32 [63] technology and provide powerful edge computing
capabilities at a low cost. This edge processing will reduce the data throughput burden on
the middleware and cloud components. With smart sensors performing feature extraction
on-chip, only the key vibration metrics will need to be transmitted instead of raw sensor
streams. Such intelligent sensors would enable highly distributed analytics, with the cloud
focus shifted to centralized data aggregation, model management, training, and decision
making. The overall architecture would become more distributed but still retain cloud
scale for global model building across bridges. Embedding intelligence directly into the
sensors will eventually allow real-time embedded inference at the true edge, while cloud
and on-premises resources will handle system-level coordination and big data analytics.
The core technologies proven here—low-cost IoT sensors, middleware processing, a
digital twin, and machine learning—provide a framework to enable low-cost, data-driven
structural health monitoring across infrastructure domains.
Author Contributions: Conceptualization, D.Z.-S. and A.A.; experimentation, A.A. and D.Z.-S.;
investigation, D.Z.-S. and A.A.; original draft preparation, A.A. and D.Z.-S.; writing—review, D.Z.-S.;
editing, D.Z.-S. and A.A.; project administration, D.Z.-S. and A.A. All authors have read and agreed
to the published version of the manuscript.
Funding: This research was funded by the Basque Government within the HAZITEK programme (ZU-
BIoT project (ZL-2020/00902)) co-financed by the European Regional Development Fund (FEDER),
and the Horizon Europe (HE) programme within MULTICLIMACT project (GA 101123538).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data are contained within the article.
Acknowledgments: The authors would like to thank Euskal Trenbide Sarea/Basque Railway Network
(ETS) and the company DAIR Ingenieros for facilitating access to the railway bridge for the pilot case
and the company RDT Ingenieros for providing the BIM model of the bridge. The authors would
also like to acknowledge Ana Fernández-Navamuel for her valuable contributions in reviewing this
article and providing key inputs and advice during the conceptual development process.
Conflicts of Interest: The authors declare there are no conflicts of interest.
Sensors 2024, 24, 2115 28 of 30
References
1. Wang, Y.W.; Ni, Y.Q.; Wang, S.M. Structural health monitoring of railway bridges using innovative sensing technologies and
machine learning algorithms: A concise review. Intell. Transp. Infrastruct. 2022, 1, liac009. [CrossRef]
2. Vagnoli, M.; Remenyte-Prescott, R.; Andrews, J. Railway Bridge Structural Health Monitoring and Fault Detection: State-of-the-
Art Methods and Future Challenges, Structural Health Monitoring. Int. J. 2017, 17. [CrossRef]
3. Elfgren, L.; Olofsson, J.; Bell, B.; Paulsson, B.; Niederleithinger, E.; Jensen, J.S.; Feltrin, G.; Täljsten, B.; Cremona, C.; Kiviluoma,
R.; et al. Sustainable Bridges—Assessment for Future Traffic Demands and Longer Lives; Priority Sixth Sustainable Development Global
Change & Ecosystems Integrated Project; Publishable Final Activity Report; Dolnoslaskie Wydawnictwo Edukacyjne: Wroclaw,
Poland, 2008.
4. Reyer, M.; Hurlebaus, S.; Mander, J.; Ozbulut, O.E. Design of a Wireless Sensor Network for Structural Health Monitoring of
Bridges. In Proceedings of the International Conference on Sensing Technology, Palmerston North, New Zealand, 28 November–1
December 2011; pp. 515–520. [CrossRef]
5. Le, B.; Andrews, J. Modelling Railway Bridge Asset Management. Proc. Inst. Mech. Eng. 2013, 227, 644–656. [CrossRef]
6. Pipinato, A.; Patton, R. Chapter 19—Railway Bridges. In Innovative Bridge Design Handbook; Butterworth Heinemann: Oxford,
UK, 2016; pp. 509–527. [CrossRef]
7. Nasr, A.; Kjellström, E.; Björnsson, I.; Honfi, D.; Ivanov, O.L.; Johansson, J. Bridges in a Changing Climate: A Study of the
Potential Impacts of Climate Change on Bridges and their Possible Adaptations. Struct. Infrastruct. Eng. 2019, 16, 738–749.
[CrossRef]
8. Rymsza, J. Causes of the Collapse of the Polcevera Viaduct in Genoa, Italy. Appl. Sci. 2021, 11, 8098. [CrossRef]
9. Wang, E.; Liu, Z.; Li, M.; Zhao, D. Implication of bridge resilience design and lessons from negative examples. Adv. Bridge Eng.
2022, 3, 23. [CrossRef]
10. Sun, L.; Shang, Z.; Xia, Y.; Bhowmick, S.; Nagarajaiah, S. Review of Bridge Structural Health Monitoring Aided by Big Data and
Artificial Intelligence: From Condition Assessment to Damage Detection. J. Struct. Eng. 2020, 146, 04020073. [CrossRef]
11. Brownjohn, J. Structural health monitoring of civil infrastructure. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2007, 365, 589–622.
[CrossRef] [PubMed]
12. He, Z.; Li, W.; Salehi, H.; Zhang, H.; Zhou, H.; Jiao, P. Integrated structural health monitoring in bridge engineering. Autom.
Constr. 2022, 136, 104168. [CrossRef]
13. Carden, E.P.; Fanning, P. Vibration Based Condition Monitoring: A Review. Struct. Health Monit. 2004, 3, 355–377. [CrossRef]
14. Alves, V.; Cury, A. A fast and efficient feature extraction methodology for structural damage localization based on raw acceleration
measurements. Struct. Control Health Monit. 2021, 28, e2748. [CrossRef]
15. Malekloo, A.; Ozer, E.; AlHamaydeh, M.; Girolami, M. Machine learning and structural health monitoring overview with
emerging technology and high-dimensional data source highlights. Struct. Health Monit. 2022, 21, 1906–1955. [CrossRef]
16. Zinno, R.; Haghshenas, S.; Guido, G.; Rashvand, K.; Vitale, A.; Sarhadi, A. The State of the Art of Artificial Intelligence Approaches
and New Technologies in Structural Health Monitoring of Bridges. Appl. Sci. 2023, 13, 97. [CrossRef]
17. Zinno, R.; Haghshenas, S.S.; Guido, G.; VItale, A. Artificial Intelligence and Structural Health Monitoring of Bridges: A Review
of the State-of-the-Art. IEEE Access 2022, 10, 88058–88078. [CrossRef]
18. Jeong, J.P.L.H.S.S.; Law, K.H. A scalable cloud-based cyberinfrastructure platform for bridge monitoring. Struct. Infrastruct. Eng.
2019, 15, 82–102. [CrossRef]
19. Vanderhorn, E.; Mahadevan, S. Digital Twin: Generalization, characterization and implementation. Decis. Support Syst. 2021,
145, 113524. [CrossRef]
20. Chiachío, M.; Megía, M.; Chiachío, J.; Fernandez, J.; Jalón, M.L. Structural digital twin framework: Formulation and technology
integration. Autom. Constr. 2022, 140, 104333. [CrossRef]
21. O’higgins, C.; Hester, D.; Mcgetrick, P.; Cross, E.J.; Ao, W.K.; Brownjohn, J. Minimal Information Data-Modelling (MID) and an
Easily Implementable Low-Cost SHM System for Use on a Short-Span Bridge. Sensors 2023, 23, 6328. [CrossRef]
22. Samudra, S.; Barbosh, M.; Sadhu, A. Machine Learning-Assisted Improved Anomaly Detection for Structural Health Monitoring.
Sensors 2023, 23, 3365. [CrossRef]
23. Anastasia, S.; García-Macías, E.; Ubertini, F.; Gattulli, V.; Ivorra, S. Damage Identification of Railway Bridges through Temporal
Autoregressive Modeling. Sensors 2023, 23, 8830. [CrossRef]
24. Sonbul, O.; Rashid, M. Towards the Structural Health Monitoring of Bridges Using Wireless Sensor Networks: A Systematic
Study. Sensors 2023, 23, 8468. [CrossRef] [PubMed]
25. Preethichandra, D.; Suntharavadivel, T.; Kalutara, P.; Piyathilaka, L.; Izhar, U. Influence of Smart Sensors on Structural Health
Monitoring Systems and Future Asset Management Practices. Sensors 2023, 23, 8279. [CrossRef]
26. Indhu, R.; Sundar, G.R.; Parveen, H.S. A Review of Machine Learning Algorithms for vibration-based SHM and vision-based
SHM. In Proceedings of the 2022 Second International Conference on Artificial Intelligence and Smart Energy (ICAIS), Coimbatore,
India, 23–25 February 2022; pp. 418–422. [CrossRef]
27. Sadhu, A.; Peplinski, J.E.; Mohammadkhorasani, A.; Moreu, F. A Review of Data Management and Visualization Techniques for
Structural Health Monitoring Using BIM and Virtual or Augmented Reality. J. Struct. Eng. 2023, 149, 03122006. [CrossRef]
28. Pierleoni, P. IoT Solution based on MQTT Protocol for Real-Time Building Monitoring. In Proceedings of the 2019 IEEE 23rd
International Symposium on Consumer Technologies (ISCT), Ancona, Italy, 19–21 June 2019; pp. 57–62. [CrossRef]
Sensors 2024, 24, 2115 29 of 30
29. Fernandez-Navamuel, A.; Zamora-Sánchez, D.; Ángel J Omella.; Pardo, D.; Garcia-Sanchez, D.; Magalhães, F. Supervised Deep
Learning with Finite Element simulations for damage identification in bridges. Eng. Struct. 2022, 257, 114016. [CrossRef]
30. Gao, Y.; Li, H.; Xiong, G.; Song, H. AIoT-informed digital twin communication for bridge maintenance. Autom. Constr. 2023,
150, 104835. [CrossRef]
31. Hu, W.; Lim, K.; Cai, Y. Digital Twin and Industry 4.0 Enablers in Building and Construction: A Survey. Buildings 2022, 12, 2004.
[CrossRef]
32. Fernandez-Navamuel, A.; Magalhães, F.; Zamora-Sánchez, D.; Omella, A.J.; Garcia-Sanchez, D.; Pardo, D. Deep learning
enhanced principal component analysis for structural health monitoring. Struct. Health Monit. 2022, 21, 1710–1722. [CrossRef]
33. Saidin, S.S.; Jamadin, A.; Kudus, S.A.; Amin, N.A.M.; Anuar, M.A. An Overview: The Application of Vibration-Based Techniques
in Bridge Structural Health Monitoring. Int. J. Concr. Struct. Mater. 2022, 16, 1–17. [CrossRef]
34. Mishra, M.; Lourenço, P.B.; Ramana, G.V. Structural health monitoring of civil engineering structures by using the internet of
things: A review. J. Build. Eng. 2022, 48, 103954. [CrossRef]
35. Zonzini, F.; Romano, F.; Carbone, A.; Zauli, M.; De Marchi, L. Enhancing Vibration-Based Structural Health Monitoring via Edge
Computing: A Tiny Machine Learning Perspective. In Proceedings of the 2021 48th Annual Review of Progress in Quantitative
Nondestructive Evaluation, Virtual, 28–30 July 2021; p. V001T07A004. [CrossRef]
36. Abner, M.; Wong, P.K.Y.; Cheng, J.C. Battery lifespan enhancement strategies for edge computing-enabled wireless Bluetooth
mesh sensor network for structural health monitoring. Autom. Constr. 2022, 140, 104355. [CrossRef]
37. Martín, C.; Garrido, D.; Llopis, L.; Rubio, B.; Díaz, M. Facilitating the monitoring and management of structural health in civil
infrastructures with an Edge/Fog/Cloud architecture. Comput. Stand. Interfaces 2022, 81, 103600. [CrossRef]
38. Sakr, M.; Sadhu, A. Visualization of structural health monitoring information using Internet-of-Things integrated with building
information modeling. J. Infrastruct. Intell. Resil. 2023, 2, 100053. [CrossRef]
39. Batchu, R.; Raisi, K.; Yu, T. Structural health monitoring of a train model under traffic loading. In Nondestructive Characterization
and Monitoring of Advanced Materials, Aerospace, Civil Infrastructure, and Transportation XVII; Shull, P.J., Gyekenyesi, A.L., Wu,
H.F., Yu, T., Eds.; International Society for Optics and Photonics, SPIE: Bellingham, WA, USA, 2023; Volume 12487, p. 124870T.
[CrossRef]
40. MLflow, Open Source Platform to Manage the ML Lifecycle. Available online: https://mlflow.org/ (accessed on 7 November 2023).
41. Ribeiro, R.R.; Lameiras, R.d.M. Evaluation of low-cost MEMS accelerometers for SHM: Frequency and damping identification of
civil structures. Lat. Am. J. Solids Struct. 2019, 16, e203. [CrossRef]
42. BeanDevice Willow AX-3D Accelerometers. Available online: https://www.beanair.com/wifi-wireless-iot-accelerometer-sensors.
html (accessed on 16 January 2024).
43. Eclipse Mosquitto, an Open Source MQTT Broker. Available online: https://mosquitto.org/ (accessed on 7 November 2023).
44. Node-Red, Low-Code Programming for Event-Driven Applications. Available online: https://nodered.org/ (accessed on
7 November 2023).
45. Nit, ulescu, I.V.; Korodi, A. Supervisory Control and Data Acquisition Approach in Node-RED: Application and Discussions. IoT
2020, 1, 76–91. [CrossRef]
46. InfluxDB Times Series Data Platform. Available online: https://www.influxdata.com/ (accessed on 7 November 2023).
47. Microsoft Azure, Cloud Computing Platform. Available online: https://azure.microsoft.com/en-us/ (accessed on 7 November 2023).
48. Introduction to Azure Data Lake Storage Gen2. Available online: https://learn.microsoft.com/en-us/azure/storage/blobs/
data-lake-storage-introduction (accessed on 7 November 2023).
49. What Is the Databricks Lakehouse? Available online: https://learn.microsoft.com/en-us/azure/databricks/lakehouse/ (ac-
cessed on 7 November 2023).
50. Databricks, Cloud Data Platform. Available online: https://www.databricks.com/ (accessed on 7 November 2023).
51. Databricks Medallion Architecture. Available online: https://www.databricks.com/glossary/medallion-architecture (accessed
on 7 November 2023).
52. AutoML, Automatically Apply Machine Learning to a Dataset. Available online: https://www.databricks.com/product/automl
(accessed on 7 November 2023).
53. Ansys Parametric Design Language. Available online: https://www.ansys.com/blog/what-is-apdl (accessed on 7 November 2023).
54. Ke, G.; Meng, Q.; Finley, T.; Wang, T.; Chen, W.; Ma, W.; Ye, Q.; Liu, T.Y. LightGBM: A Highly Efficient Gradient Boosting
Decision Tree. In Advances in Neural Information Processing Systems; Guyon, I., Luxburg, U.V., Bengio, S., Wallach, H., Fergus, R.,
Vishwanathan, S., Garnett, R., Eds.; Curran Associates, Inc.: Red Hook, NY, USA, 2017; Volume 30.
55. Noori Hoshyar, A.; Rashidi, M.; Yu, Y.; Samali, B. Proposed Machine Learning Techniques for Bridge Structural Health Monitoring:
A Laboratory Study. Remote Sens. 2023, 15, 1984. [CrossRef]
56. Eldeib, A.; Abdelsalam, A.; Shehata, A.; Ali, H.; Fouad, S. Health monitoring of historic buildings using machine learning in
real-time internet of things. Indones. J. Electr. Eng. Comput. Sci. 2023, 32, 725. [CrossRef]
57. Noel, A.B.; Abdaoui, A.; Elfouly, T.; Ahmed, M.H.; Badawy, A.; Shehata, M.S. Structural Health Monitoring Using Wireless Sensor
Networks: A Comprehensive Survey. IEEE Commun. Surv. Tutorials 2017, 19, 1403–1423. [CrossRef]
58. Chintalapudi, K.; Fu, T.; Paek, J.; Kothari, N.; Rangwala, S.; Caffrey, J.; Govindan, R.; Johnson, E.; Masri, S. Monitoring civil
structures with a wireless sensor network. IEEE Internet Comput. 2006, 10, 26–34. [CrossRef]
Sensors 2024, 24, 2115 30 of 30
59. Liu, X.; Cao, J. Structural Health Monitoring Using Wireless Sensor Networks. In Mobile and Pervasive Computing in Construction;
Anumba, C.J., Wang, X., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2012. [CrossRef]
60. Noghabi, S.A.; Kolb, J.; Bodik, P.; Cuervo, E. Unified Management and Optimization of Edge-Cloud IoT Applications. arXiv 2018,
arXiv:1805.02305. https://doi.org/10.48550/arXiv.1805.02305.
61. Energy Harvesting Ambient Energy Managers. Available online: https://e-peas.com/energy-harvesting/ (accessed on
6 March 2024).
62. Voltaic Systems. Available online: https://voltaicsystems.com/ (accessed on 6 March 2024).
63. Espressif Systems. Available online: https://www.espressif.com/en/products/socs/esp32 (accessed on 6 March 2024).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.