Sensors 22 04610 v2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

sensors

Review
LiDAR-Based Structural Health Monitoring: Applications in
Civil Infrastructure Systems
Elise Kaartinen, Kyle Dunphy and Ayan Sadhu *

Department of Civil and Environmental Engineering, Western University, London, ON N6A 3K7, Canada;
[email protected] (E.K.); [email protected] (K.D.)
* Correspondence: [email protected]

Abstract: As innovative technologies emerge, extensive research has been undertaken to develop new
structural health monitoring procedures. The current methods, involving on-site visual inspections,
have proven to be costly, time-consuming, labor-intensive, and highly subjective for assessing the
safety and integrity of civil infrastructures. Mobile and stationary LiDAR (Light Detection and
Ranging) devices have significant potential for damage detection, as the scans provide detailed
geometric information about the structures being evaluated. This paper reviews the recent develop-
ments for LiDAR-based structural health monitoring, in particular, for detecting cracks, deformation,
defects, or changes to structures over time. In this regard, mobile laser scanning (MLS) and terrestrial
laser scanning (TLS), specific to structural health monitoring, were reviewed for a wide range of
civil infrastructure systems, including bridges, roads and pavements, tunnels and arch structures,
post-disaster reconnaissance, historical and heritage structures, roofs, and retaining walls. Finally,
the existing limitations and future research directions of LiDAR technology for structural health
monitoring are discussed in detail.

Keywords: terrestrial laser scanning; mobile laser scanning; structural assessment; automation;
Citation: Kaartinen, E.; Dunphy, K.;
damage detection; quality control
Sadhu, A. LiDAR-Based Structural
Health Monitoring: Applications in
Civil Infrastructure Systems. Sensors
2022, 22, 4610. https://doi.org/
10.3390/s22124610 1. Introduction

Academic Editors: Maurizio


Infrastructure, including roads, bridges, and buildings, has significantly deteriorated
Spadavecchia, Nicola Giaquinto
on a global scale, as many of these structures have a limited duration of exploitation.
and Francesco Adamo
The prevalence of deteriorating structures, in combination with the increased frequency
of natural disaster events within recent decades, has necessitated the development of
Received: 9 May 2022 inspection techniques to ascertain the damage and overall level of safety of structures.
Accepted: 16 June 2022 Often the nature of these structural damages results in logistic constraints, by which human-
Published: 18 June 2022
based visual inspections cannot be conducted, as they pose a danger to the inspectors’
Publisher’s Note: MDPI stays neutral health or the area under inspection cannot be accessed by normal means. Additionally,
with regard to jurisdictional claims in the use of contact-based devices is time-consuming as they require an extensive setup to
published maps and institutional affil- attach the devices to the structure and initiate the data collection process. This ultimately
iations. delays the rehabilitation process for damaged structures, often resulting in economic losses.
Therefore, the need for rapid inspection methods which implement non-contact sensors
within the field of Structural Health Monitoring (SHM) has become more prevalent within
the past decades.
Copyright: © 2022 by the authors.
A non-contact sensor [1] can be defined as any sensor which is implemented for
Licensee MDPI, Basel, Switzerland.
monitoring and data acquisition that does not require the sensor to have physical contact
This article is an open access article
with the structures. Broadly, these devices can be classified into two categories of data
distributed under the terms and
collection techniques, based on the nature of data collection: (1) passive and (2) active.
conditions of the Creative Commons
Passive techniques use optical setup to take images or videos of the damaged structures,
Attribution (CC BY) license (https://
which can later be analyzed by specialists to determine the level of damage present and
creativecommons.org/licenses/by/
4.0/).
the necessary rehabilitation technique required. Active techniques emit various spectral

Sensors 2022, 22, 4610. https://doi.org/10.3390/s22124610 https://www.mdpi.com/journal/sensors


Sensors 2022, 22, 4610 2 of 32

bands and frequencies of light and sound which are directed at the structure. The device
then measures the Doppler shifting of the returning sound wave or reflected light to
ascertain the nature of anomalies in the object under investigation. They address the
limitations of human-based inspections as they allow for inspections to be conducted
without physical intervention, and their generally efficient setup allows for inspections to
be conducted quickly.
The growing popularity of non-contact sensors with SHM has resulted in cameras,
smartphones, unmanned aerial vehicles, satellites, ultrasonic devices, and various other
sensors being used to conduct structural inspections [1]. Optical-based sensors, such as
cameras, smartphones, unmanned aerial vehicles, and satellites, capture an image or video-
based data, which can be used to assess the overall stability of a structure. The data is
often directly stored on the device that captures the image-based information or can be
transmitted back to a base station for further analysis in the case of mobile optical sensors.
The quality of data extracted from these devices is highly susceptible to environmental
factors, such as lighting, wind, and mechanical vibrations due to moving machine parts [1,2].
Moreover, markers must often be affixed to the structure to quantify any physical change in
the structure during monitoring, including displacement or strain. Alternatively, acoustic-
based sensors, such as ultrasonic devices, detect damages in structural elements, such as
cracks and delamination, through the variation of ultrasonic waves as they propagate
through the structural mass. However, non-contact ultrasonic devices are limited to
analyzing the surface of the structure, which only provides details on surficial damages.
Moreover, these types of devices have numerous variants which have limited applications
to a specific type of monitoring, making it difficult to implement in a generalized setting [2].
Light Detection and Ranging (LiDAR) devices and methods have received significant
consideration in the past decade by the SHM research community as a tool for 3D point
cloud generation and analysis of structures. Recently, many studies have proposed method-
ologies to optimize the collection and processing of data to create more time-efficient, and
computationally inexpensive, SHM systems. An unmanned aerial vehicle-based LiDAR
with an optimized flight path decision-making paradigm was implemented by [3] for the
inspection of surface defects for bridges. Furthermore, state-of-the-art algorithms have
been proposed for the detection of structural elements and damages within millimeters of
precision, while minimizing error [4,5]. Over the past decade, several literature reviews
have been conducted by researchers pertaining to the implementation of LiDAR devices for
structural health monitoring. However, the majority of published literature review papers
broadly assess non-contact sensing techniques for damage detection, rather than focus-
ing on LiDAR devices, specifically [6–12]. Furthermore, current LiDAR-based literature
review papers focus on very limited domains within SHM, including deformation monitor-
ing of buildings [13], bridge inspection [14,15], and moisture detection in structures [16].
Therefore, the objectives of the proposed review summarized in this paper are as follows:
1. Provide a comprehensive review of LiDAR-based SHM assessment for various aging
civil infrastructure systems, including bridges, tunnels, roads, roofs, walls, buildings,
and historical structures.
2. Summarize the existing limitations and gap areas for LiDAR-based SHM technology.
3. Articulate the future research direction of the LiDAR-based monitoring and inspection
techniques for SHM.

2. Background of LiDAR
Applications of LiDAR for SHM typically involve analyzing 3D objects, structural ge-
ometries, deformations, crack information, and visualization using a dense 3D dataset with
finer resolution and precision [6–16]. In general, the applications are of two types: (a) phase-
based (rapid but limited to short distances) and (b) time-of-flight-based (able to measure
large distances, but slow and less accurate). Various laser scanners are currently available
with a range of speeds, typically 2000–120,000 points per second, maximum resolutions
typically 1–100 mm at 50 m, and accuracies typically 3–50 mm at 100 m [10–13,17–19].
phase-based (rapid but limited to short distances) and (b) time-of-flight-based (able to
measure large distances, but slow and less accurate). Various laser scanners are currently
Sensors 2022, 22, 4610 available with a range of speeds, typically 2000–120,000 points per second, maximum 3 of 32

resolutions typically 1–100 mm at 50 m, and accuracies typically 3–50 mm at 100 m [10–


13,17–19]. They can be used for as-built building information modeling, structural
They can beand
inspection, used for as-built building
reconnaissance surveys,information
by capturingmodeling,
3D data and structural inspection,
meta-data and
of the target
reconnaissance
several hundredsurveys,
meters by capturing
away 3D dataofand
to a precision a fewmeta-data of the target several hundred
millimeters.
meters away to a precision of a few millimeters.
2.1. Advantages and Limitations of LiDAR Sensors
2.1. Advantages and Limitations of LiDAR Sensors
A potential advantage of LiDAR devices over traditional non-contact sensors is that
thereAispotential advantage
no requirement for of LiDARinstrumentation
physical devices over traditional non-contact
and accessibility sensors
of the is that
structures,
there is no
thereby requirement
saving for physical
a significant amount of instrumentation
time and labor and costsaccessibility
and improving of the thestructures,
safety of
thereby saving a significant amount of time and labor costs and improving
the workplace. Moreover, these devices are independent of natural light sources, the safety of the
serving
workplace. Moreover, these devices are independent of natural light
as excellent field measurement tools. However, LiDAR has several following limitationssources, serving as
excellent field measurement tools. However, LiDAR has several following
that prompted SHM researchers to advance this technology over the last several years. limitations that
prompted SHM researchers to advance this technology over the last several years.
1. Dependency on weather conditions.
1. Dependency
2. Expensive, onperformance
and weather conditions.
degrades with distance.
2. Expensive, and performance
3. Non-structural components must degrades
often with distance.using specific filters.
be removed
3. Non-structural components must often be
4. The datasets captured by LiDAR devices are extremely removed using specific filters. that the
large, meaning
4. computational
The datasets captured by LiDAR devices
processing time is long. are extremely large, meaning that the com-
putational processing time is long.
5. Strategic positioning is required to avoid overlapping or targetless scans. The field of
5. view,
Strategic positioning
or the is required
angle covered by thetosensor,
avoid overlapping
impacts whatoristargetless
capturedscans.
by theThe field of
sensor.
view, or the angle covered by the sensor, impacts what is captured by
6. There is often the need for multiple scans at different points; thus, object extraction the sensor.
6. There is often the need for multiple scans at different points; thus, object extraction or
or interpolation techniques are required.
interpolation techniques are required.
2.2.
2.2. Types of LiDAR
Types of LiDAR Sensors
Sensors
Depending
Depending onon the mounting mechanism,
the mounting mechanism, their
their application
application can
can be categorized into
be categorized into
two classes, based on mode of operation: (a) stationary and (b) mobile.
two classes, based on mode of operation: (a) stationary and (b) mobile.

2.2.1.
2.2.1. TLS
TLS
Terrestrial laser scanning (TLS) is a stationary instrument (as shown in Figure 1) that
can acquire
acquire dense
densepoint
pointclouds,
clouds,often
oftentaken
taken from
from various
various locations
locations close
close to the
to the ground.
ground. For
For
SHM, SHM,
TLSTLS
scansscans are captured
are captured and and
mergedmerged to create
to create larger
larger pointpoint clouds.
clouds. The The
size size of
of this
this
pointpoint cloud
cloud depends
depends on whether
on whether the scanner
the scanner is short-range,
is short-range, medium-range,
medium-range, or long-
or long-range.
range. Although
Although usingoften
using TLS TLS requires
often requires
more more
manual manual
laborlabor and time,
and time, thesethese point
point clouds
clouds are
are generally
generally moremore detailed
detailed andand precise
precise thanthan mobile
mobile laserlaser scanning
scanning (MLS).
(MLS).

Figure
Figure 1.
1. A
A typical
typical TLS:
TLS: (a)
(a) FARO Focus 130
FARO Focus 130 3D
3D laser
laser scanner;
scanner; (b)
(b) Leica
Leica TC2002
TC2002 total
total station
station [20].
[20].

2.2.2. ALS
Mobile Laser Scanning (MLS) refers to a broad category of devices/instruments that
are mounted on mobile equipment, such as drones, airplanes, cars, or helicopters, as shown
without this manual operation. ALS is like TLS in the sense that the device/instrument
also emits laser signals and calculates the distance, based on the time delay of the returned
signals.
The difference is that the mobile LiDAR device can travel up to 100 km/h and is free
to rotate, which requires the need for a GPS receiver and an inertial measurement unit
Sensors 2022, 22, 4610 (IMU). The GPS allows the airborne device/instrument to record the exact location of 4 ofthe
32
system to estimate where the device/instrument reflections are located on the scanned
surfaces. The distance recorded by each returned laser pulse is subtracted from the
recorded
in Figure 2.altitude of the
Airborne system
Laser to yield
Scanning theiselevation
(ALS) of the
particularly scanned
useful surface. point
for acquiring To account
cloud
for the
data of tilt
largeof areas
the system,
of landwhen
(e.g.,the distance
towns, of the returned
neighborhoods, pulse is
or cities), calculated,
surfaces the hard
that are IMU
records
to accessthe roll,skyscrapers
(e.g., yaw, and pitch of the system
or bridges), and atforeach location. Most situations
time-constrained pulses are (e.g.,
emitted at
post-
differentsituations).
disaster angles, soTLSthe surveys
system typically
also accounts
comprisefor the pulsescans
multiple anglefrom
when calculating
multiple the
positions;
elevation ALS
however, of each point.capture
surveys This information
point clouds onatthe speed,heights
different angle, and
and rotation
distancesofwithout
the mobile
this
manual
system operation. ALS is like
are all important forTLS in the sense
accurate that the
elevation device/instrument
calculations also emits
and, therefore, forlaser
the
signals and calculates
acquirement the distance,
of an accurate based on the time delay of the returned signals.
point cloud.

Figure 2. A typical ALS integrated into a drone [21].

The
Table difference
1 shows is that the mobile
a detailed LiDARofdevice
comparison TLS andcanALS
travelin up to 100
light km/h
of their and is free to
applications in
rotate, which requires the need for a GPS receiver and an inertial measurement unit (IMU).
SHM [13–19].
The GPS allows the airborne device/instrument to record the exact location of the system
to estimate
Table where the
1. Comparison device/instrument
of advantages reflections
and disadvantages are and
of TLS located
ALS.on the scanned surfaces.
The distance recorded by each returned laser pulse is subtracted from the recorded altitude
TLS of the scanned surface. To account
of the system to yield the elevation ALS for the tilt of the
• More detailed and precise •
system, when the distance of the returned pulse is calculated, the IMUautomated
ALS is highly records the androll,
point cloud. requires less manual movement
yaw, and pitch of the system at each location. Most pulses are emitted at different angles, of
Advantages • Better control over the the instrument.
so the system also accounts for the pulse angle when calculating the elevation of each
point. This informationcaptured
on thepoint
speed,cloud; and• rotation
angle, less It requires
of theless emphasis
mobile systemonare each
all
redundant data to filter out scan’s angle and target,
important for accurate elevation calculations and, therefore, for the acquirement of an as it can be
accurate point cloud.
Table 1 shows a detailed comparison of TLS and ALS in light of their applications in
SHM [13–19].

Table 1. Comparison of advantages and disadvantages of TLS and ALS.

TLS ALS

• ALS is highly automated and requires less manual


• More detailed and precise point cloud. movement of the instrument.
• Better control over the captured point cloud; • It requires less emphasis on each scan’s angle and
less redundant data to filter out during the target, as it can be maneuvered at different heights
Advantages post-processing stages. off the ground, which captures more points.
• TLS-based technology is less complex to • More points are captured than in typical TLS scans,
operate than ALS, which incorporates GPS and therefore, larger surface areas can be covered by
IMU elements. the scans.
• It can capture harder-to-reach targets (e.g., roofs of
buildings) and larger areas in a shorter period.
targets (e.g., roofs of build
and larger areas in a sh
period.
• TLS is highly manual and
Sensors 2022, 22, 4610 requires the movement of the• Pre-planned flight5 ofpaths32 are
scanner to capture the required to reduce the amou
intended target(s). redundant point cloud data.
Table 1. Cont.
• There is less flexibility in the• These devices are hi
TLS device’s mobility; thus, ALSdependent on weather condit
incomplete or obstructed due to consistent speeds b
• Disadvantages
TLS is highly manual and requires the
movement of the scanner to capture scans
the are common.
• Pre-planned flight pathsrequired for uniform
are often required to point clo
intended target(s). • Larger point reduce clouds take• of redundant
the amount With pointan cloud
abundance
data. of
• There is less flexibility in the device’s mobility;
longer •
amounts These
of devices
time are
to highlycaptured,
dependent on weather
ALS point cl
Disadvantages thus, incomplete or obstructed scans conditions, due to consistent speeds being required
are common. acquire. for uniform point clouds.require more complex and t
• Larger point clouds take longer • amounts
TLS devices
of • often
Withare unable of data
an abundance consuming
captured, ALS pointpost-proces
time to acquire. clouds require
to reach remote or hard-to- more complex and time-consuming
methods.
• TLS devices often are unable to reach remote or post-processing methods.
hard-to-reach locations. reach locations.

3.3.Literature
Literature Review
Review of LiDAR-Based
of LiDAR-Based SHM SHM
Figure
Figure 3 summarizes
3 summarizes various
various applications
applications of ALS
of ALS and TLS and TLS in
in a broad a broad
range range of S
of SHM
applications in aging civil infrastructure.
applications in aging civil infrastructure.

Figure3. 3.
Figure Applications
Applications of LiDAR
of LiDAR for SHM.
for SHM.

3.1. Applications in Large-Scale Civil Infrastructure


3.1.1. Road/Pavement
Mobility issues also limit the uses by which laser-based scanning can be conducted
for pavement damage detection. The sensor must be directly attached to a vehicle to make
the data acquisition process time-efficient. Ref. [8] reviewed the current methods and tech-
nologies for detecting damages (i.e., cracking, patching, potholes, surface deformation, and
surface defects) on pavements. It was concluded that distresses that manifest primarily in
height and depth differences require 3-D sensors to detect information in three dimensions.
Sensors 2022, 22, 4610 6 of 32

Ref. [22] developed a computational process to semi-automatically extract the irregulari-


ties of a concrete taxiway, based on DEM and TLS data. Once the DEM was constructed
using an interpolation algorithm, the elevation profiles and the values of the longitudinal
grade and transverse slopes were extracted. In another study, ref. [23] concrete slabs were
evaluated, by calculating the faulting values between them using a TLS. With reference to
the general plane, the altitude and the distance between the points of each slab-plane were
computed to identify the critical sections. For each of these sections, the faulting values
were determined by computing the distance along the normal to the general plane.
Ref. [24] identified the potential causes of longitudinal cracking in Jointed Plain
Concrete Pavement using TLS. Curling and warping were measured by the TLS at various
sites, and the average curvature degree of the slabs was further computed. Ref. [25]
proposed a data processing framework for identifying distress (e.g., potholes, swells, and
shoves) in the pavement from MLS data. After the points belonging to the road surface were
extracted, the height deviation of each point was computed for a modeled road surface.
Two binary images were created from the DEM to analyze the positive and negative
displacements. Finally, each segmented region of the full-scale pavement was classified,
based on its severity levels. Ref. [26] used TLS data to scan a large concrete area to evaluate
defects in the rigid pavement slabs. The acquired point cloud was cut into nine sub-clouds
to represent each rigid slab, and the defects were characterized, based on three parameters
(i.e., defected area, crack width, and intensity).

3.1.2. Bridges
LiDAR-based SHM assessment of bridges has predominately involved characterizing
the geometric properties of various components (girders, deck, etc.). This information
is often used to assess the vertical clearance underneath bridges to ensure that large
trucks can safely pass beneath the structure. Various studies using LiDAR data have been
conducted to ascertain the correlation between environmental properties, such as loading,
temperature, and precipitation, on bridge deformation. A few studies have implemented
LiDAR for the quantification of various structural damages, including spalling and crack
detection. For example, ref. [27] proposed an automated method to recognize the mass
loss of concrete bridges using TLS. The damage recognition was performed in three steps:
(1) the point clouds were subdivided into sub-areas with a defined size, (2) a preliminary
Gaussian filtering and parabolic fitting was performed for each point in each sub-area, and
(3) each sub-area was classified as damaged or undamaged based on the corresponding
curvature distribution.
A LiDAR-based bridge evaluation for material mass loss quantification was explored
by [28]. In a bridge-monitoring study, a phase-based laser system was used to compare both
the distance and gradient-based damage quantification methods to detect the defected area.
The results showed that combining the two approaches improves the identification and
quantification capability of the LiDAR. In another study, ref. [29] developed an automatic
and high-precision bridge clearance measurement technique based on TLS. The relevant
part of the bridge was selected from the planar scan image to calculate the clearance,
showing millimeter-level accuracy. In a similar direction, ref. [30] evaluated the impact
of various parameters (i.e., elastic shortening, creep, shrinkage, relaxation, and thermal
expansion) on bridge clearance measurements in a full-scale structure using periodic TLS
scans. The applications, reliability, and evaluation methodologies of LiDAR technology
for bridge health monitoring were assessed by [31]. Three sensitivity analyses (i.e., linear
dimensional analysis, surface area analysis, and volumetric analysis) were performed on a
full-scale bridge to test the system adjusting parameters, the range measurement ability,
and accuracy of the scanner, as well as the automatic bridge inspection algorithms.
A novel procedure was presented [32] to measure the minimum vertical under clear-
ance of bridges and to obtain the profile of prestressed concrete beams using photogram-
metry and TLS surveys. To estimate the vertical under clearance and prestressed concrete
beam cambers, a 3D curve-fitting algorithm was developed. After implementation on a
Sensors 2022, 22, 4610 7 of 32

full-scale bridge, high statistical correlation coefficients were achieved. Three case studies
presented by [14] were used to validate the use of TLS scans for identifying various damage
characteristics, such as loss of concrete, reinforcement corrosion, and surface erosion, on
full-scale bridges. Under various truck loadings, the geometrical information (i.e., bridge
elevation, span length, girder spacing, bottom flange width, and web height) was recorded
by a LiDAR in [33]. The girder deflections were calculated by comparing the girder eleva-
tion coordinates of the scans with and without the weight of the trucks. In comparison to
contact methods, the proposed approach proved to be superior in accuracy and accessibility,
even in estimating the bridge’s natural frequencies. A fully automated TLS point cloud
segmentation procedure was developed by [34] for the SHM of masonry arch bridges. A
voxelization process first filtered out the redundant data from the point cloud, from which
segmentation was performed by combining a heuristic method with image processing. To
identify the individual structural elements, topological constraints were used to establish
the spatial relation and order of the elements. Five data sets were used to test the proposed
algorithm, showing coherent results with minor problems due to poorer point cloud quality.
Terrestrial photogrammetry was used by [35] to detect cracks in masonry arch bridges,
while ground-penetrating radar was used to detect the hidden bridge elements. In combin-
ing this data, continuum damage and discrete models were constructed to determine the
behavior of each arch under various loading situations. Two loading tests conducted on
bridges by [36] were used to determine their vertical deformations using TLS. Rather than
through individual points, the deformation analysis was conducted by one modeled surface
serving as the reference surface. After the surface meshes were created, the zero-load epoch
mesh was subtracted from each load epoch mesh. A new method called MCrack-TLS
was established by [37], which combined image-processing procedures and TLS to assess
3D crack characteristics (i.e., width, length, orientation, and location) in concrete bridges.
The TLS was used to identify and record reference target coordinates, removing the need
for real coordinates of at least four reference targets required for traditional methods. A
homography matrix, along with the acquired reference targets, was used to compute the
image orthorectification. Testing on a concrete viaduct showed that the proposed method
increased productivity and the quality of data compared to traditional methods.
An automated processing method of laser scanning data for the SHM of piers in
masonry arch bridges was proposed by [38]. A full-scale bridge was first segmented, based
on its structural elements (cutwaters, piers, spandrel walls, roadway, and vaults), so that
the data size was reduced and each pier’s face could be analyzed individually. Structural
faults were identified from the geometric parameters of the pier faces and their respective
topological relation to the other bridge elements. SHM technology for a masonry bridge,
using TLS data and Ground-Penetrating Radar (GPR), was presented by [39]. Comparing
the TLS measurements with the historical drawings and plans of the bridge, the anomalies
identified in the hyperbolic reflections provided information on the configuration of the
bridge fillings. A structural deformation analysis framework was presented by [40] using
TLS for applications on composite footbridges. Two mesh models were generated using
the Fast Marching algorithm to check if a change occurred after a load was applied to the
bridge. A crack identification and quantification approach for concrete structures were
investigated by [41] using unmanned aerial vehicles. To establish an inspection map, a point
cloud-based background model of the structure was first generated through a preliminary
flight. From high-resolution images, regions with convolutional neutral networks for deep
learning were applied for crack detection. A combination of 384 crack images and a model
pre-trained through transfer learning were used for classification and localization for a
full-scale bridge.
A deflection and deformation measurement application for bridges was proposed
by [42]. The shape information model was first constructed using the improved octree
data structure and the TLS data. This process reduced the size of the scan data for efficient
memory usage and then estimated the deflection, based on the octree space division, which
was then validated using LVDT. A new bridge inspection technique using UAV (Unmanned
Sensors 2022, 22, 4610 8 of 32

Aerial Vehicle) imagery point clouds was developed by [43]. To reduce errors in data
(i.e., incomplete data, nonuniform distribution, outlier, surface deviation, and geometric
accuracy), a triangular mesh and density map were constructed. The iterative closest
point algorithm was applied with TLS data, and thickness, point distribution, and point-
to-point distances were measured for a full-scale bridge. A reflector-based framework
by [44] was applied for measuring long-term bridge displacement using LiDAR. The
framework consisted of a reflector positioning strategy, measuring the reflector coordinates,
and calculating the displacement. Having the reflectors as reference points allowed higher
measurement accuracy, reduced scanning time, and there was no requirement to keep the
same positioning of the LiDAR at different epochs. A prestressed concrete bridge was used
to validate the accuracy of the proposed method. A TLS was used to construct a Digital
Surface Model and identify the potential damage area in [45]. Ground-based microwave
interferometry was applied to confirm where the bridge had damage, and an interferometry
synthetic aperture radar technique was applied to analyze the causes of the damage of a
full-scale bridge.
A three-dimensional path planning method for LiDAR-equipped UAVs was intro-
duced by [3] for inspecting bridges. The proposed method consists of three steps, (1)
assigning low, medium, and high Important Values, based on moment and shear force
values obtained from structural analysis, (2) selecting View Points of Interest for perpen-
dicular and overlapping views, and (3) calculating the optimal collision-free path using
Genetic Algorithm and A* algorithm. After being implemented on a full-scale bridge, it
was concluded that the proposed path planning method decreased flight time, processing
time, and workload while increasing visibility, reliability, and accuracy. A deformation
monitoring process, based on TLS and ground-based radar interferometry data, was es-
tablished and the process tested on a full-scale bridge by [46]. Using the TLS data from
three epochs, the bridge’s vertical displacements were determined using a geometry-based
approach, while another deformation was computed-based on deviation from the reference
points. Comparison between the ground-based radar interferometry and the TLS showed
close compliance with the results from both methods. In another recent study, ref. [47]
proposed a method to obtain a displacement estimation of bridge structures using four
laser scanning-based techniques. The vertical displacement was estimated by relocating
the point cloud data in a 3D space and then dividing it in detail to search for the change in
position of the leaf nodes. This study rearranged the points in a three-dimensional space,
and nodes were created to calculate the displacement. Comparing the Grid, Tri, and LSP
approaches of using distance estimation between points, the proposed method showed
a decrease in the time required for displacement estimation, but an increase in the data
processing time.
To date, the use of TLS for quality inspection has been focused on identifying surface
defects and the presence of water in structural members through utilizing differential
geometry, RGB values, image segmenting, and gradient-based methods, as well as combin-
ing different image-based technologies, as summarized by [15]. For assessing structural
performance, the common methods have been to construct geometric models as the ba-
sis for measuring deformations. The use of TLS and photogrammetric techniques for
measuring the point-wise aspects (distances and lengths) was implemented by [20] for a
historic suspension bridge. It was concluded that the distance from the bridge, as well as
the complexity of the bridge, both highly influenced the accuracy of the measurements.
Moreover, the hybrid surveying method acquired millimeter-level accuracy measurements,
but the TLS performed better than the photogrammetric device. The B-spline surface
method for the approximation, deformation analysis, and noise filtering of point clouds
using TLS was quantified by [48]. After testing on a bridge under load, it was concluded
that the mathematical approximation of the noisy point cloud was necessary for accurate
computation, and the correlated noise impacted the distance computation for both the
raw and approximated observations. A plane fitting approach was used in [5] to classify
the points into six sub-plane categories (e.g., the bottom of bottom flange, edge of bottom
Sensors 2022, 22, 4610 9 of 32

flange, front edge effect plane, side of the web, bottom of top flange, and back edge effect
plane) and fitted the points to the corresponding dataset using a linear least square method.
Finally, Table 2 summarizes the devices used, type of assessment, and post-processing
method for the LiDAR-based structural assessment of bridges.

Table 2. A comprehensive summary of laser-based assessment techniques for bridges.

References Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


Important value analysis with
[3] N/A Full-scale inspection
genetic and A* algorithms
The plane fitting
[5] FARO Focus 3D Geometric assessment
least-squares method
Material mass loss, erosion,
[14] N/A DGC and MMSET methods
and corrosion
[15] N/A Full-scale inspection
[20] Leica TC2002 Geometric assessment Hybrid photogrammetric methods
[27] Riegl LMS Z-420i Material mass loss Curvature distribution
[28] FARO LS 880HE Material mass loss Distance and gradient-based
LiDAR bridge evaluation
[29] FARO LS 880HE Bridge clearance
(LiBE) method
Impact of parameters on
[30] N/A Correlation analysis
bridge clearance
Sensitivity analysis and
[31] N/A Bridge health monitoring
LiBE method
[32] Riegl LMS Z-390i Bridge clearance 3D curve-fitting algorithm
Geometrical assessment under
[33] FARO LS 880HE The difference in elevation data
various loadings
Voxelization and
[34] Riegl LMS Z-390i Full-scale inspection
topological constraints
Continuum damage and
[35] Leica TCR1102 Geometric reconstruction
discrete models
[36] Leica ScanStation C10 Structural deformation The difference in surface profiles
[37] Riegl LMS Z-390i 3D crack characterization MCrack-TLS
[38] Riegl LMS Z-390i Pier analysis Geometric and topological analysis
Comparison of scanned data
[39] FARO Focus 3D X330 Full-scale inspection
and drawings
[40] Leica ScanStation C10 Structural deformation Fast marching algorithm
Region-based Convolutional
[41] N/A Crack identification
neural network
Octree space partitioning
[42] Leica ScanStation C5 Structural deformation
(OSP) algorithm
[43] Leica ScanStation P20 Full-scale inspection Closest point algorithm
[44] Riegel VZ-1000 Structural deformation Reflector coordinates analysis
[45] Riegel VZ-1000 Damage detection Digitial surface model analysis
[46] Leica ScanStation2 Structural deformation Geometry-based analysis
Various displacement estimate
[47] N/A Structural deformation
methods and OSP
Hausdorff distance and averaged
[48] Zoller + Frohlich Imager 5006H Structural deformation
derivation comparison
Sensors 2022, 22, 4610 10 of 32

3.1.3. Tunnels
For underground tunnels, excessive-profile deformations are a significant concern as
they often result in the collapse of the structure. Therefore, the majority of LiDAR or TLS-
based inspections of these tunnels focus on ongoing data acquisition, which can be used to
measure the change in profile deflection with respect to time. An approach for monitoring
tunnel profile deformations with the use of multi-epoch LiDAR was established by [49].
The method was based on establishing point correspondences between the point clouds
from different epochs and applying a minimum-distance projection algorithm to identify
deformations. An autonomous technique for extracting tunnel cross-sections and removing
non-lining points, based on TLS point clouds, was proposed by [50]. The first step was
to estimate the tunnel boundary points using an angle threshold and 2D projection onto
the X-Y plane. The direction of the cross-sectional plane was adjusted twice with the total
least-squares method and Rodrigues’ rotation formula. Finally, an angle-based filtering
algorithm removed non-lining points, based on morphological erosion. Validation on a real
railway tunnel indicated that the proposed method was superior in accuracy compared
to other methods, due to the consideration of the tunnel grade and the application of
the filtering algorithm. A method to calculate the clearance of tunnels, based on mobile
laser scanning, was developed by [51]. For the pre-processing step, the point cloud was
segmented in the direction of the rail line, and the straight rail sections were identified.
Based on the tunnel cross-section baseline and the individual cross-sections, the clearance
inspection was carried out by calculating the distance between the cross-section and the
testing rack.
A method for the automatic extraction of tunnel cross-sections was investigated by [52]
based on mobile laser scanning to monitor deformations of a full-scale tunnel. First, the 3D
point clouds were converted to a 2D surface, where the buffer of each cross-section was
calculated by the K-Nearest neighbor algorithm. The iterative ellipse fitting was applied by
combining the fitting of the sectional curve line with the denoising of the sectional point
set. To reduce errors, the initial cross-sectional planes were rotated around their respective
intercept points, and the optimal cross-sections were extracted. In [53], a processing method
based on TLS for the change detection of the cross-sectional area of an underground gate
road was explored. Three reference points were first established to reduce errors when data
were compared from different scanner positions. A visualization software package was
developed to visualize and analyze the deformations from different epochs. In another
study, ref. [54] investigated the feature extraction of tunnel structures using TLS technology.
The method comprised of segmenting the tunnel’s point cloud data into thin sections to
acquire the projection plane for each section profile. Millimeter-level accuracy was achieved
when the deflection of a subway tunnel was estimated.
A new method was established by [55] that considered and corrected the effect of
surface roughness on TLS intensity data for water leakage detection in underground tunnels.
After the mean intensity values of each homogeneous region were computed, the distance
and incident angle effects were corrected to improve the accuracy of the intensity data.
The intensity image of the studied tunnel was generated, and the water leakage regions
were detected based on an intensity threshold. Crack identification in tunnels by using TLS
data was further explored in a study by [56]. After the point cloud was projected onto an
image, an index method was used to indicate the position of each crack. To extract each
crack, the standard deviation of the Gaussian template was used as a parameter, and a
signal-to-noise ratio further extracted the smaller cracks. A crack detection method for
tunnels was presented by [57], combining dilation and the Canny algorithm based on
TLS point cloud data. The grayscale dilation process was employed to eliminate distinct
textures in each image by using a disk-shaped structuring element. The Canny detector was
adopted for crack edge detection, where the vertical crack widths were measured, based on
the space of the two sharp peaks. The proposed method did not require determining Canny
detector parameters, thereby being free of any major user intervention. Table 3 summarizes
Sensors 2022, 22, 4610 11 of 32

the devices used, type of assessment, and post-processing method for the LiDAR-based
structural assessment of bridges.

Table 3. A comprehensive summary of LiDAR-based SHM for tunnels.

Reference Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


[49] Zoller + Frohlich 5010 TLS Profile deformations Minimum-distance projection algorithm
Least squares, Rodrigues’ rotation and
[50] FARO X130 Tunnel cross-section
angle-based filtering
[51] N/A Tunnel clearance Differences between Sequential Profiles
K-Nearest Neighbor and iterative ellipse
[52] Regel VUX-IHA Profile deformations
fitting Algorithm
[53] N/A Tunnel cross-section Differences between profiles
[54] N/A Profile deformations Circular filtering and RANSAC algorithm
[55] Riegl VZ-400i Water leakage detection Intensity thresholding method
[56] N/A Crack identification Index and Gaussian template methods
[57] N/A Crack identification Dilation and canny algorithm

3.2. Applications in Civil Structural Sytems


3.2.1. Arched Structures
Similar to underground tunnels, excessive-profile deformations in arched structures
are a significant concern for the continued integrity of the structure and must be monitored
through TLS devices. An experiment was carried out by [58] on a brick and concrete
arch structure under monotonic loads, based on TLS measurements. A surface model
was created, and the deformation of the arch was calculated by subtraction of two epoch
surfaces. The result was a sufficiently accurate dense point-wise representation of the
deformation. A new TLS-based method, as proposed by [59], extracted the displacement
size and direction of arched structures. A network and remapped point cloud were used
to take the structural feature points and compare the exact points between two epochs.
Analysis conducted by [60] quantified the deformation tendency of a composite masonry
structure, based on TLS data extraction by the window selection method. The window-
neighbor method first extracted the edge data. Moreover, polynomial fitting was applied to
analyze and compare the deformation between each epoch during a loading test of an arch
structure. Approximating TLS point cloud data for deformation analysis was conducted
by [61] for arch structures through polynomial and B-splines surface models. To select an
optimal parametric model, different adjusted surface models were compared through Cox’s
and Vuong’s tests. It was concluded that the B-spline model was superior and performed
better with more parameters.
A novel algorithm developed by [62] was implemented for the extractions of steel
arches from a LiDAR point cloud of varying tunnel cross-sections (e.g., round and square).
To extract the rock portion of the point cloud, slices were made along the X-axis, and the
Differential Analysis for the Section Sequences of the Tunnel point cloud was applied to
create curvature and height thresholds. Finally, the normal local saliency of each point
served as the basis for extracting the steel arches from the rock surface. Results from testing
on a full-scale tunnel showed the proposed method’s ability to extract the steel arches at
92.1% precision without manual assistance. TLS and the finite element method (FEM) were
combined in [63] to construct an intelligent FEM model that could predict and simulate the
deformation of arched structures. An experiment was conducted where a concrete arch was
measured and inspected during different loading scenarios, and the resulting deformations
and patterns were recognized to create an optimized FEM model. Table 4 summarizes
the devices used, type of assessment, and post-processing method for the LiDAR-based
structural assessment of arched structures.
Sensors 2022, 22, 4610 12 of 32

Table 4. A comprehensive summary of LiDAR-based SHM for arched structures.

Reference Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


Zoller + Frohlich Imager 5006 and
[58] Profile deformations Differences between sequential profiles
Leice Laser
[59] Zoller + Frohlick Imager 5006 Profile deformations Differences between sequential profiles
Window-neighbor method and
[60] Zoller + Frohlick Imager 5006 Profile deformations
polynomial fitting
[61] Leica AT960LR Profile deformations Polynomial and b-splines models
Differential analysis and normal
[62] P + F R2000 UHD Tunnel cross-section
local saliency
[63] N/A Profile deformations Finite element methods

3.2.2. Historical/Heritage Structures


Preservation and rehabilitation of historic structures have become a key component of
SHM, as many structures globally have reached, and exceeded, their service lives. Historic
structures are often characterized by extreme fragility, and as such often require techniques
that reduce physical interaction with the structure. Therefore, the use of non-destructive
testing methods, such as laser-based scanning, allows inspectors to assess the structure
without direct physical contact with the structure. A comparison of three methods (i.e.,
isodata algorithm, k-means algorithm, and fuzzy k-means algorithm) for characterizing
historical building damage, using the intensity data from three different TLSs, found that
the fuzzy k-means algorithm resulted in the best accuracy [64]. Evaluating the geometric
characteristics (i.e., the global and local leaning and tapering angle, radius, local deviation
from the circular shape, and the local curvatures) and the damage characteristics (i.e.,
masonry bulges, brick displacements, material loss, and cracks) of a historic masonry tower
by [65] was conducted using TLS data. A toolbox was developed for the analysis of the
point cloud, which provided an accurate georeferenced model of the structure. A case study
on the SHM of a historical masonry building produced by [66] combined TLS and infrared
(IR) thermal images for vulnerability analysis. TLS data were used for the 3D reconstruction
of the structure, while the anomalies were detected from IR thermography. Metric features
were assigned to the thermographic images, which mapped the anomalies and made them
measurable. Finally, the vulnerability of the structural elements was assessed, based on the
thermal defect being associated with geometric irregularity.
The vertical deformation of a minaret of a historical mosque was studied by [67]
using TLS. Thirty-one horizontal sections of the minaret point cloud were created, which
determined the deformation for each element from the calculated inclinations. Detailed
deformation analysis of an ancient wooden structure was conducted by [68], combining
three TLSs with different distances (long-range, middle-range, and handheld). After adding
the coordinates of the target points into the coordinate system, the points from the three
different scan locations were registered and geo-referenced into one dataset to perform the
deformation analysis. Moisture content was quantified in [69] for heritage buildings based
on autonomous TLS data acquisition. Once the point clouds were acquired, they were
processed to the optimal level of reflectivity through an algorithm that determined if the
incidence angle was within a certain threshold. The range of reflectivity was divided into a
set of linear segments to create a model showing the moisture and humidity content, based
on the change in the reflectivity index. Using TLS data, in a study by [70], a watertight mesh
model was created and imported into the FE software to perform structural analyses. The
deviation analyses used the point-cloud-based mesh model, to identify local deformations.
The combination of the FE model and DA provided accurate and non-destructive gathering
of information for structural analysis and monitoring. A numerical modeling strategy
proposed by [71] implemented TLS point clouds for building FE models and SHM of
historical buildings. The structural breakdown allowed a semi-automatic generation of the
Sensors 2022, 22, 4610 13 of 32

structural domain from the TLS. The validation of the strategy was performed on a historic
fortress, which demonstrated increased level of automation and decreased computational
time, compared to CAD-based modeling procedures.
Mobile LiDAR Systems were evaluated by [72] for the analysis of cultural heritage
sites, based on a two-fold approach. A clustering phase consisted of computing the local
curvature, defining the number of clusters, according to similar curvature values, and
conducting a component analysis to reduce errors. Lastly, a weighted sampling was
applied to each point inside a cluster, based on the extreme curvature values of the cluster
and the associated feature (e.g., cylinder, plane, etc.). Testing on a medieval wall and an
accuracy comparison using TLS showed reduced data acquisition time, but also reduced
spatial resolution of the mobile LiDAR point cloud. A method to detect and localize
deformation, specific to historic and heritage buildings, was proposed by [73], which used
a generalized Procrustes analysis (GPA) and TLS data. The acquired data set was divided
into subsets, and the GPA was applied by forming a matrix containing three-dimensional
coordinates of each point of the point cloud. The deformation vectors and probability were
computed, based on six transformation parameters. Various testing methods in historic
sites showed the ability of the proposed method to reduce noise and improve the reliability
and accuracy of the results. In another study, ref. [74] investigated the use of TLS for the
post-fire inspection of a historic building. TLS scans were taken of the building before
and immediately after the fire, and cloud-to-cloud registration techniques were used to
identify the changes in the common features. The progressive decay and erosion of earthen
heritage sites were assessed by [75] by combining multi-temporal TLS data and GIS. Using
the Multiscale approach to the Model Cloud Comparison method, the surface change
among different instances of the same feature was computed. Full-scale heritage walls
were evaluated, and the deterioration values were imported into the GIS to express the
occurrence of the variation.
The use of multi-temporal TLS data comparisons was explored in [76], based on the
Multiscale Model to Model Cloud Comparison (M3C2) technique to detect material loss
in ancient walls and buildings. Each instance of a compared feature was aligned to its
reference point cloud, and the point normals were computed to detect change. The M3C2
method compared a sub-set of points based on a cylindrical projection from user-defined
maximum depth and radius. Analysis of various damage assessment methods by [77]
for heritage building elements was conducted, based on TLS. For the inner and outer
walls, three different methods were tested: (1) approximating the point cloud centered on a
point, (2) considering the plane as vertical and moving the origin point of that plane, and
(3) considering a set of points within a certain distance to calculate the vertical plane. An
automated deep learning model was developed by [78] for the surface damage detection of
heritage sites using 2D images and 3D point clouds. After the data acquisition, Semantic
Segmentation was applied to the images to remove the sources of noise and to generate
a per-pixel classification of each image. The proposed method was validated after being
tested on an unseen heritage site, proving accurate damage detection of complex heritage
structures. Through a voxelating process, the point-to-point spacing was made uniform,
such that the damaged areas of the point cloud showed a different point distribution [79].
Based on the eigenvalues, neighboring points covariance matrix, normal vector variation,
and mean curvature to its closest neighboring points of each point, the damage was detected
and re-evaluated. From the damaged areas of the point cloud, a density-based clustering
algorithm was applied, which identified clustering structures for the categorization of the
damage. Table 5 summarizes the devices used, type of assessment, and post-processing
method for the laser-based inspection of historic structures.
Sensors 2022, 22, 4610 14 of 32

Table 5. A comprehensive summary of laser-based assessment techniques for historic structures.

References Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


Isodata algorithm, k-means
FARO Photon, TRIMBLE GX200
[64] Damage analysis algorithm, and fuzzy
and Rieg Z-390i
k-means algorithm
[65] ILRIS 3D Geometric and Damage Analysis MATLAB Octave Toolbox
[66] Zoller + Frohlick Imager 5010c Damage analysis Thermography and 3d model analysis
[67] Trimble GX 200 Deformation analysis Inclination calculations
Riegl VZ 1000, FARO Focus 3D
[68] Deformation analysis Digital reconstruction
and Handyscan 3D
[69] Leica HDS-3000 Moisture measurement Reflectivity-based model
Finite element model and
[70] FARO Focus 3D S-120 Deformation analysis
deviation analysis
[71] N/A Damage analysis Finite element model
[72] LYNX Mobile Mapper Damage analysis Clustering and weighted sampling
Leica ScanStation P20, Leica T830, Probability analysis of
[73] Deformation analysis
and Leica P40 ScanStation deformation vectors
Leica ScanStation 2, Leica
ScanStation C10, FARO Focus 3D Cloud-to-cloud
[74] Damage analysis
120, and Zoller + Frohlick Registration technique
Imager 5010c
[75] N/A Damage analysis Multiscale model cloud comparison
Multiscale model to model
[76] FARO Focus S120 Damage analysis
cloud comparison
[77] Leica RTC360 Damage analysis Point cloud analysis
Region-based
[78] N/A Damage analysis
Convolutionalneural network
[79] FARO Focus 3D X120 Damage analysis Density-based clustering algorithm

3.2.3. Concrete Structures


Laser-based scanning of concrete structures provides engineers in the SHM commu-
nity with information about the severity of the damage. Furthermore, the 3D point clouds
generated by these techniques provide quantifiable data about the level of cracking and
spalling, which is easily obtainable through other optic-based devices, such as cameras
or traditional surveying equipment. Three algorithms proposed and tested by [80] (i.e.,
range filtering, deviation filtering, and sliding window) were implemented for processing
laser-scanned data for the assessment of surface flatness deviations. The range filtering and
deviation filtering algorithms relied on range images with reversed orders of smoothing
and deviation calculations, while the sliding window algorithm operated on 3D points to
determine the deviation at each surface location. Testing on flat boards with defects of vary-
ing diameters and thicknesses showed that the sliding window algorithm offered the best
detection performance. Two methods for quantitatively evaluating the damage in concrete
structures, based on long-distance TLS, were proposed by [81]. Based on the region grow-
ing algorithm, the first method estimated the original shape of a structure before scaling
and evaluated the total scaling depth. The second method was proposed using the iterative
closest point algorithm, combined with a novel feature sampling technique, to evaluate
secular changes in the scaling depth. An autonomous method by [82] was implemented
for detecting and mapping cracks in concrete obtained from laser scanning surveying. To
remove the noise elements, different filtering and image processing procedures were first
executed. The probabilistic relaxation technique was applied to extract the crack track from
Sensors 2022, 22, 4610 15 of 32

the filtered image, and the pixel coordinates were transformed into global 3D coordinates.
For mapping the crack, the results from the experiment showed an accuracy of 10–38 mm,
compared to the total station survey.
Comparisons were made by [83] between surface-based TLS measurements and the
FEM model simulation of the displacement and loading of concrete cylinders. An error
of less than 5% proved the feasibility of using TLS for the evaluation of FEM models. A
novel approach developed by [84] integrated TLS with Building Information Modelling
(BIM) to assess the flatness of concrete surfaces. The data processing was significantly
automated by using the Scan-vs-BIM method, which matched the point cloud to the BIM
model components. The Straightedge and F-Numbers methods were applied to control
the compliance of the surfaces. The proposed method was successfully applied, as per the
current standards for flatness specification and control in concrete slabs. An unexplored
and novel technique to simultaneously localize and quantify spalling defects on concrete
surfaces using a terrestrial laser scanner was conceived by [85]. Two defect-sensitive
features (i.e., angle and distance deviation) were combined to identify defects, while a
defect classifier was developed to diagnose the severity, location, and size of defects. A
suite of scan parameters (e.g., scan distance, angular resolution, incidental angle) was
investigated in the parametric study and numerical stimulations. The results showed
that the proposed technique improves the autonomy, simultaneousness, and accuracy of
detecting concrete defects.
A two-stage research project on reinforced concrete structures was proposed by [86]
using laser scanning technology for crack identification and monitoring. Although the
TLS data provided information on the swelling and surface height change of the concrete
blocks, it was not possible to measure the cracks, due to the low scanning resolution and
small crack widths. A surface normal-based damage detection method, as developed
by [87], was implemented to detect and quantify damage types, such as cracks, spalling,
corrosion, delamination, and rupture, using camera-integrated TLS. The defective areas
were located using the model properties and grouped into individual damage clusters
using a silhouette-based method. The quantitative information of the damage clusters was
recorded using the presented damage area and volume computation strategies. The results
from testing on a full-scale concrete test frame and a bridge showed that the proposed
method automatically quantified and documented information, thus eliminating the need
for human and computer interaction. TLS data was explored by [88] to identify and quantify
areas of concrete loss in structures. The proposed approach was validated using accelerated
laboratory testing to determine the feasibility of TLS for identifying crack initiation and
subsequent crack growth, followed by crack monitoring in a forty-year-old reinforced
concrete seawall.
A novel concrete crack detection method proposed by [89] applied an adaptive wavelet
neutral network for the analysis of TLS data. A low-resolution fit of the entire 3D point
cloud data was created, and further high-resolution analysis was performed only on regions
with damage. Such an approach resulted in a compact representation of the TLS data,
thus reducing memory usage and computational time. The detection of cracks on concrete
structures by using image processing algorithms from the octree structure of TLS data
was explored by [90]. Testing on a concrete dam showed that the proposed technique
minimized the false recognition of cracks against stains, sediment, and structural joints. A
concrete surface crack detection method by [91] combined the use of 2D images and 3D
point clouds. The depth information from the laser scanners and gray information were
merged at the pixel level. The improved Otsu algorithm yielded rough crack detection;
therefore, denoising and connection of the cracks were required to refine the results. After
several types of cracked concrete specimens were tested, the approach showed significantly
better results than the single image or standalone point cloud methods. Integrating UAV-
equipped LiDAR data and RGB images, ref. [92] identified and quantified cracks in concrete.
The points within the point cloud were first associated with their corresponding structural
element using the heuristic-based method. Once the crack patches were identified through
Sensors 2022, 22, 4610 16 of 32

a CNN-based classifier, an adaptive thresholding procedure of the grayscale intensity


values aided in extracting the pixels with crack boundaries. Table 6 summarizes the devices
used, type of assessment, and post-processing method for the laser-based inspection of
concrete structures.

Table 6. A comprehensive summary of laser-based assessment techniques for concrete structures.

References Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


[80] N/A Surface roughness Filtering and sliding window
[81] Riegl VZ-400 Scaling detection Region growing algorithm
[82] Leica C10 TLS Crack detection Probabilistic relaxation technique
[83] N/A Displacement measurement Finite element model
[84] FARO Focus 3D Surface roughness Scan-vs.-bim method
Angle and distance deviation
[85] FARO Focus 3D Spalling detection
with classifier
[86] Trimble CX Crack detection Point cloud analysis
Crack, Spalling, Corrosion,
[87] FARO Focus 3D and Photon 80 Delamination and Silhouette-based Method
Rupture Detection
Multiscale model to model
[88] Trimble TX5 3D Spaling detection
cloud comparison
[89] Trimble TX5 3D Crack detection Low and high-resolution fit
K-means clustering, median
[90] Leica ScanStation C5 Crack detection
filtering, and otsu’s binarization
[91] Reigl VZ-2000 and HandySCAN 700 Crack detection Otsu’s binarization
[92] Velodyne VLP-16 Crack detection Convolutional neural network

3.2.4. Retaining Walls


Very few studies have been conducted regarding the inspection of retaining walls
for damage defects using a laser-based scanning device. An automated framework was
introduced by [93] for the feature extraction and displacement measurement of highway
retaining walls using TLS data. The method was based on extracting the horizontal
joints from the wall’s point cloud and using this data as a benchmark for detecting future
displacements. Testing on a real-life dataset and 3D simulated models showed millimeter-
level accuracy of the displacement measurements. A time-efficient method was developed
by [94], intended for the morphological characterization of masonry blocks of a medieval
wall, using TLS and MLS data. The 3D point cloud was reduced to 2D intensity images by
computing the plane of projection and converting the 2D point cloud into a raster image.
The watershed segmentation process involved differentiating the masonry blocks from the
joints on the basis that the gray tone of each pixel represented the height of the surface. A
change detection method, based on comparing baselines (a 3D line segment connecting two
feature points in one scan) from the laser scanner data of two epochs, was developed by [95].
The target points (i.e., spherical targets, planar targets, or virtual points) were first extracted
utilizing k-means clustering of the TLS intensity data. Comparing the two baselines from
the different scans connecting the same features identified the changes in the x, y, and z
directions. The proposed framework was validated during experiments on masonry walls,
which showed reduced errors, without the need for a registration step.
Continuous Wavelet Transform (CWT) was applied to the 2.5D map using an estimate
of the mortar joint width in [96]. The resulting scalogram showed the CWT responses
for each pixel in the depth map. Furthermore, the dilation process delivered properly
defined 2D stone segments to be re-mapped onto the 3D point cloud. The proposed method
was tested on two structures for the evaluation of the changes in the individual stones,
Sensors 2022, 22, 4610 17 of 32

showing no sensitivity to global levels of flatness, waviness, curvature, and plumpness


of walls. Rough surfaces were shown to have lower intensity values in [16], due to the
laser signal being reflected several times. Also, the color of a surface affected reflectivity,
with darker surfaces having an increased absorption of light and decreased energy of the
return signal. A targetless and MLS-based SHM strategy for mechanically stabilized earth
(MSE) walls was developed by [97]. The MSE wall façade was partitioned into individual
planar faces, and the longitudinal and transversal lines along each face were established.
The serviceability measures were determined based on the translational and rotational
relationships, as well as the normal displacement between the corners of the fitted planes
and each panel. The proposed method was validated when the MLS results from a segment
of a full-scale MSE wall were compared with TLS and a profiler gauge.
A framework to reduce the size of point cloud datasets, while also detecting sur-
face imperfections (i.e., cracks and cavities) and physicochemical issues (i.e., moisture,
weathering, salt blooming, and biodeterioration), in building walls was explored by [98].
The point clouds were downsized using geometric and intensity data, both of which had
thresholds associated with undamaged surfaces. From the remaining point cloud, the two
types of data were analyzed, and defects were identified. A methodology for monitoring
and detecting defects in levees was developed by [99], based on multi-temporal LiDAR
point clouds and RGB images. The LiDAR data was used to produce two differential
Digital Terrain Models, and the elevation changes were identified. The RGB images and
calculated vegetation indices were used to search for changes in the levee land cover. A
method for measuring the tilt and lateral displacement of retaining walls using mobile
laser scanning was introduced by [100]. From the acquired point cloud, the retaining
wall was extracted from the ground points through a binary classification. As opposed
to computing point-to-point deviations, the anchored concrete panels were individually
segmented and modeled with planar surfaces. Furthermore, changes in the plane’s key
parameters indicated deformations between two epochs. Table 7 summarizes the devices
used, type of assessment, and post-processing method for the LiDAR-based inspection of
retaining walls.

Table 7. A comprehensive summary of laser-based assessment techniques for retaining walls.

Reference Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


[16] N/A Moisture detection Reflectivity analysis
[93] Riegl VZ-400 Displacement measurement Difference analysis
Raster image and
[94] Riegl LMS Z-390i Morphologic characterization
watershed segmentation
K-means clustering and
[95] Leica ScanStation C10 Change detection
difference analysis
Continuous wavelet
[96] Leica P40 and FARO Focus 3D Wall segmentation
transform and dilation process
[97] Riegl VUX-1HA and ZF Profile 9012 Displacement measurement
[98] Trimble TX8 and Leica C10 Defect detection Point cloud analysis
[99] GNSS Leica GS15 and CS15 Defect detection Dsm analysis
[100] Zoller + Frohlick 9012 Displacement measurement

3.2.5. Roofs
Applications for laser-based inspection of roofs have been limited within the SHM
field, in comparison to other analyses involving the entire 3D structure or various elements
of the structures. As unmanned aerial vehicles with a sufficient payload capable of carrying
a laser-based scanning device are required to conduct the survey, the relative accessibility
of rooftop inspections becomes restricted. A segmentation method using laser data for
Sensors 2022, 22, 4610 18 of 32

damaged roofs was proposed by [101]. The individual points were grouped into planar
regions, based on the assumption that undamaged roofs appeared as planar segments and
collapsed roofs comprised of many small segments. The roofs were identified as intact or
damaged based on the extracted features and a classifier trained from manually labeled
segments. The automatic extraction of roofs through a data-driven approach by [102]
integrated LiDAR data and multispectral ortho-imagery for improved city modeling and
building inspection. The LiDAR data was divided into ground and non-ground points,
where the non-ground points were further segmented to extract the roof planes. Using
the ground mask, color, and textural information, the structural image lines were put into
various classes (i.e., ground, tree, roof edge, and roof ridge). Various algorithms were
applied to obtain a roof plane and remove planes constructed on trees. The proposed
method was tested on two data sets and successfully extracted small planes and removed
vegetation. An automated roof covering damage assessment method by [103], based
on ground LiDAR, collected data in the aftermath of extreme winds. Experiments were
conducted in a controlled laboratory where the k-means clustering algorithm was tested
with different combinations of clustering evaluation criteria.
The structural damage of roofs was detected in [104], based on their 3D features
extracted from only the post-event airborne LiDAR data. During the data pre-processing,
the Digital Surface Model (DSM) for each building was created using post-earthquake
LiDAR data, 2D GIS vector data, and the digital elevation model. For each building,
damaged roofs were detected based on the 3D shape descriptor derived from the contour
clusters of the DSM. The proposed method was validated using post-earthquake data,
proving the better performance of the 3D shape descriptor compared to using geometric
features for damage detection. A robust methodology to evaluate tornado fragility models
with roof damage was conducted by [105] using post-tornado LiDAR data. The extracted
geometric information (i.e., height, slope, pressure zones, and distance to tornado path)
for each roof was used to produce fragility curves for roof sheathing failures. Based on the
LiDAR data, the tornado wind speeds and building damage were calculated and compared
with the values estimated by the fragility curves.

3.2.6. Post-Disaster Reconnaissance


Natural disasters, such as earthquakes, landslides, hurricanes, and floods, often create
logistical constraints for structural inspection, thus increasing the difficulty in performing
human-based inspections. As such, non-contact devices that can be deployed at a distance
from the damaged structure, such as LiDAR devices, are increasingly being implemented
for post-disaster reconnaissance of structures and hazard regions. Two processing strategies
for post-event ALS data were compared by [106] to automatically detect collapsed build-
ings. A segmentation algorithm first identified the planar regions, and the geometric and
radiometric attributes were calculated. The rule-based classifications strategy identified
collapsed buildings, based on each segment’s attributes and the corresponding thresholds.
The Maxent classifier detected collapsed buildings, based on the maximum entropy model-
ing. Both strategies were relatively accurate, although the rule-based classifier required
some manual input of the threshold values, and the Maxent classifier was highly dependent
on the availability of precise training data. The advantages and limits of using remote
sensing for post-earthquake damage extraction were summarized by [6]. They concluded
that the open issues were the definition of a damage scale based on the data, the use of
data fusion techniques, and the use of crowd mapping procedures. A TLS-based method to
rapidly evaluate earthquake-induced damages to buildings was proposed by [107]. The
point clouds were first interpolated to create the primitive plane. The computation of
the differences between the primitives and the points produced morphological maps. If
multi-temporal data was available, difference maps, as well as the detections of changes of
the primitives, were obtained. The proposed procedure was validated by three case stud-
ies being conducted on earthquake-damaged structures, all with varying multi-temporal
data available.
Sensors 2022, 22, 4610 19 of 32

An evaluation of multi-temporal and mono-temporal methods for identifying earthquake-


induced building damage from optical, LiDAR, and SAR data was reviewed by [7]. The
common technique using pre- and post-event LiDAR data involves comparing multi-
temporal 3D building models to assess the basis of the damage classification. With only
post-event data available, algorithms have been developed for automatic plane detection
from LiDAR data. A mobile LiDAR was employed by [108] for post-disaster data collection
and to develop mapping approaches for damage assessment. From the data collection after
a hurricane, the laser scans and imagery were tied together, employing post-processing of
the trajectory of the vehicle. The change was detected by comparing the pre- and post-event
data, and it was concluded that the proposed approach yielded more detailed information
and analysis than traditional methods. The use of laser scanning technology for civil
engineering laboratory tests and reconnaissance of earthquake-damaged structures was
discussed by [109]. A method for the delineation of earthquake-damaged buildings from
an image-based 3D point cloud (Blom-CGR) was developed by [110] and identified the
broken elements of the buildings that led to the gaps in the point cloud. The gaps were
then classified into four categories (i.e., occlusion, failure in 3D point generation, opening
in architectural design, and damage) based on their surrounding damage patterns.
A framework was introduced by [111] that estimated the damage of structural frame
members using information from post-earthquake point clouds and the pre-damaged
BIM model. The scanner’s point clouds were registered in the local cadastral coordinate
system and segmented to construct planar surfaces. Objects from the as-built BIM model
were compared to the segmented point cloud, and any deviations (i.e., cracks or breaks)
were identified and updated in the as-damaged BIM model of RC frames. The Structure
from motion (SFM)-based dense reconstruction method for conducting a post-hurricane
residential building damage assessment was investigated by [112]. The generated 3D point
clouds obtained from mobile LiDAR data were compared with SURE and Autodesk’s
123D Catch. A quantitative approach by [113] implemented airborne LiDAR data for the
identification of post-earthquake building damage for buildings with different roof types.
To extract the severity of building damage, the study utilized surface normal algorithms
and the ratio of the standard deviation to the mean absolute deviation of the angle between
the surface normal and the zenith. The authors concluded that the proposed method
effectively estimated damage independent of the roof style and without the need for pre-
earthquake data. Similarly, ref. [114] presented an automated method for the segmentation
and damage assessment of post-earthquake buildings using airborne LiDAR obtained from
1953 buildings.
A damage evaluation framework by [115] integrated LiDAR scan data and photogram-
metry technologies. The damaged estimation was enhanced by a joint analysis being
conducted on the results from both the image-based method and the 3D coordinate method.
A study using LiDAR scan data of an earthquake-damaged wall of a building proved that
the proposed method significantly improved the accuracy of the damage assessment results.
The use of LiDAR data to detect damaged buildings by using digital surface models (DSMs)
before and after an earthquake was investigated by [116]. A total of 26,128 buildings were
evaluated based on three parameters: the average change in height, its standard deviation,
and the correlation coefficient between the two DSMs. Results showed that the most in-
fluential factor for damage detection was the change in average elevation. An automated
approach for the post-disaster structural damage evaluation of major building envelope
elements (i.e., wall, roof, balcony, column, and handrail) was developed by [117] using
mobile LiDAR data. The building was first semantically parsed into segments, and damage
detection was conducted to extract the semantic structural damage information. The study
was the first automated building component-level damage assessment with high-resolution
point cloud data sets.
A novel airborne LiDAR-based approach by [118] was implemented for the assessment
of a wide scale of post-hurricane building damage. The building objects were first extracted
from pre-event and post-event data, and a cluster-matching algorithm was used to compute
Sensors 2022, 22, 4610 20 of 32

the differences between the extracted building objects from the multiple-temporal data
sets. On a hierarchical basis, the damage was estimated based on the damage indicators.
Compared to other methods, the proposed approach effectively-recognized building objects
extracted damage features, and characterized the extent of damage all at the individual
building level. A comprehensive framework by [119] detected various damage types
(i.e., multilayer collapse, outspread multilayer, pancake collapse, upper stories collapse,
heaps of debris, collapse of all floors, inclined plane buildings, and inclined to overturn
collapse) based on post-event LiDAR data. One conventional and two novel texture
extraction strategies were used to generate the textural features. An improved Vosselman
filtering method identified pancaked buildings, and the inclination angles were estimated
from LiDAR data. From ambient vibration results analyzed in [120], modal parameters
were extracted, and an FE model was subsequently created and updated. The LiDAR
data was used to quantify the defects and to compare with the estimated damage from
the vibration measurements, which showed good agreement with the developed model.
Table 8 summarizes the devices used, type of assessment, and post-processing method for
laser-based post-disaster reconnaissance.

Table 8. A comprehensive summary of LiDAR-based assessment techniques for post-disaster reconnaissance.

References Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


Pre- and
[7] N/A Earthquake-induced building damages
post-Event comparison
Segmentation algorithm with
[106] Lecia ALS50 Collapsed building detection maxent and
rule-based classifiers
[107] N/A Earthquake-induced building damages Difference maps
Pre- and
[108] Optech LYNX Mobile Mapper M1 Hazard maps-hurricanes
post-event comparison
Leica ScanStation C10 and
[109] Earthquake-induced building damages N/A
ScanStation 2
Gabor Wavelets, Support
[110] Blom-CGR Earthquake-induced building damages Vector Machines and
Random Forest
Point Cloud and BIM
[111] N/A Earthquake-induced building damages
model comparison
Pre- and
[112] Optech LYNX Mobile Mapper M1 Hurricane-induced building damages
post-event comparison
Surface Normal Algorithms
[113] N/A Earthquake-induced building damages
and Standard Deviation Ratio
[114] Lecia ALS60 Earthquake-induced building damages Classification algorithm
Hybrid method-3d coordinate
[115] FARO Focus 3D Damage evaluation
and image-based
Correlation coefficients
[116] Leica ALS50II Earthquake-induced building damages
of dsms
[117] Optech LYNX Mobile Mapper M1 Structural damage evaluation Segmentation technique
Clustering
[118] N/A Hurricane-induced building damages
matching algorithm
[119] Leica ALS60 Damage evaluation Vosselman filtering method
[120] N/A Earthquake-induced building damages Modal analysis
Sensors 2022, 22, 4610 21 of 32

3.2.7. Other Structural Members


A recent study by [121] consisted of deriving beam-deflection equations using beam
mechanics based on two TLSs. Two experiments on concrete and timber beams were
analyzed using the least-squares solutions before and after the statistical testing. The
modeling of raw point data for surface representation was undermined by the modeled
TLS data yielding significantly higher accuracy than the TLS’s coordinate precision. A
displacement measurement model was presented by [122] for SHM using TLS. After the
shape information was acquired from the TLS, the base vectors were generated using the
least-square method. The vectors were used to transform from the TLS coordinate system
to the structural coordinate system. The displacements of a steel beam were computed
by acquiring the deformed shape information, transforming the coordinates, and using
the least square method. The proposed model improved measurement accuracy, while the
maximum deflection was ~1.6% of that obtained from traditional LVDTs.
The feasibility of TLS to perform structural change and deformation analysis was
outlined in a study by [123], which proposed a method to apply TLS in a large-scale
experimental setting. During the analysis of the point clouds, the specimen was sliced
by the element, and the cross-sections were used to determine the volumetric change
(i.e., bulging, compression, and spalling). In comparison to conventional means, the
presented approach allowed for increased visualization of the specimen change and faster
identification of damage. A study on two 2D reinforced concrete frames by [124] tested
the use of TLS for measuring structural deformations under lateral loading. The acquired
point clouds were first filtered and converted into 3D coordinates. Applying the root mean
square errors of transformation to the coordinate differences identified the deformed points.
A review of recent methods for change detection was conducted by [125] using mobile
and static laser scanning data. The main challenges that were identified were point cloud
registration, varying measurement geometry, varying positions of data acquisition, and
temporary objects present in scans. Furthermore, it was emphasized that the signal-to-
noise ratio should be computed to evaluate the redundancy of the point clouds in the
processing steps. A novel first-ever algorithm by [126] was designed to automatically
measure the deformation from torsion and the deflection from the bending of metal beams
by using TLS, as well as to determine the location of maximum stress. The process consisted
of segmenting the beam flange and using polynomial surface fitting to eliminate noise
and errors. The planar surface was fitted to the point cloud, where the deformations
and stresses were calculated using the LiDAR data. The proposed methodology was
tested on a steel beam under various loading situations, proving that it yielded accurate
results. An algorithm proposed by [127] used TLS data that measured and modeled the
deformation of metal beams using segmented point clouds and the polynomial fitting of
the deformed curved surfaces. A three-stage process model for the deformation analysis
of structures was developed by [13], based on a review of recent TLS-based methods. The
change detection methods have, thus far, primarily involved computing and comparing
the distances between point clouds or fitted surfaces of two epochs (i.e., point-to-point,
point-to-surface, and surface-to-surface).
Alpha-Shapes were applied in [128] to determine the crack outlines in timber beams,
and a minimum area was fixed to delete any false cracks. The geometric characteristics of
each crack (i.e., main direction, area, crack centroid, length, and maximum width) were
obtained for future comparisons and monitoring. Testing on laboratory specimens and
a timber roof validated the suitability of the proposed algorithm for finding cracks and
detecting growth between epochs. The deformation behavior of arch and beam structures
under static loads, based on TLS and the surface analysis method, was explored by [129].
The scan data from each structural element was used to create a polynomial fit of the
beam and a surface fit of the arch to be compared. The displacement was calculated from
the deviations between the scans of different epochs. A 3D point cloud change analysis
approach was presented by [130] for detecting changes in structural inspections. Once
two 3D point clouds were spatially registered, comparisons were performed using the
Sensors 2022, 22, 4610 22 of 32

point-wise distance estimations to track and quantify displacements on a per-point basis.


To overcome distorted measurements, a statistical sampling technique was used to extract
a measurement from a localized set of per-point measurements. A series of flexural tests
were performed to evaluate the proposed approach, showing that it is an accurate and
automatic way to track small movements of structures over time. Methods were explored
in [131] for providing quantitative estimations of textural damage (i.e., tile spall off, metal
rusting, and water staining) using LiDAR data. The intensity and RGB model information
were first used for the clustering analysis, and textural damages were detected by applying
four data clustering algorithms (i.e., k-means, fuzzy c-means, subtractive clustering, and
density-based spatial clustering). The results showed that the RGB information was not
effective in detecting textural damage and that the k-means and fuzzy c-means algorithms
gave better clustering performance and computational efficiency.
A novel 3D surface descriptor was created by [132] for the automatic identification
and classification of surface defects using point cloud data from a 3D reconstruction system.
The local models on each 3D surface were first estimated and then used to calculate the
differences between the model’s normals in a local region. The primitives were projected
onto a plane, where the defects were recognized from the geometrical features of the 2D im-
ages. Using a support vector machine, the defects were classified into three categories (i.e.,
holes, bumps, and cracks) with improved robustness to noise. Laser scanner data was used
to generate a 3D Surface mesh in [133] for welding inspections, and the photogrammetric
data generated a 3D point cloud both for quality assessment purposes. It was concluded
that the accuracy of both technologies was similar; however, the macro-photogrammetric
technique yielded superior geometric and radiometric resolution. A method proposed
by [134] was applied to measure the strain and deformation of a steel plate using discretized
3D coordinate data of a LiDAR. The deformation shape of the plate was first modeled
by a high-order polynomial function through regression analysis. Using a finite element
method, a strain measurement model was generated, from which the strain on the steel
plate was estimated.
A method for the inspection of piping components was created by [135], based on
comparing the as-built TLS and the as-designed CAD data. The geometric parameters were
recognized from the acquired point cloud via a normal-based region growing segmentation
and the RANSAC method. Performing geometric parameter comparison and distance-
based deviation analysis on each pipe segment identified differences from the as-designed
model. The use of TLS for identifying and measuring surface imperfections and distor-
tions of precast concrete elements was evaluated by [136]. The proposed techniques for
evaluation were validated through precast concrete bridge deck panels. An approach for
the Finite Element (FE) model updating damaged structures was developed by [137] and
used to compare the results to LiDAR data. A two-story masonry building was measured
at its reference state, where both tuned and un-tuned FE models were created to test the
effects of the material properties and existing damage on updating FE models. After forced
vibration tests were performed with four exterior walls being removed, the final FE models
showed good accuracy, compared to the LiDAR results.
The use of a TLS, compared to Close-Range Photogrammetry (CRP) based on the
Structure from Motion (SfM) algorithm, was evaluated by [138] for measuring structural
deformations. A case study was conducted on two reinforced concrete beams subjected to
four-point bending loading conditions. The point clouds at various loading stages were
compared by using the Mesh to mesh and modeling with geometric primitives methods.
The modeling approach yielded better results than the Mesh to mesh, while the CRP results
were more accurate than the TLS, due to the short capturing distance and dimensional
scale of the images. Using the RGB values of the TLS data, 2D images were created and
segmented for precast concrete elements (PCEs) in [139]. Based on the edge image, the
active window method was applied, which extracted the important data within each image
cluster. The RBNN algorithm was used to avoid under-segmentation, and the segments
were reconstructed to avoid over-segmentation. Compared to other methods, the proposed
Sensors 2022, 22, 4610 23 of 32

technique improved the data acquisition efficiency for the quality inspection of PCEs.
An automated framework for the extraction of structural components (i.e., column, slab,
and rebar) from point cloud data for progress monitoring and compliance control during
construction was proposed by [140]. From the registered point clouds, the planar and linear
features were extracted and semantically labelled into various categories. In comparing
the as-built and the planned BIM, deviations were identified and visualized. Five sets of
TLS point clouds from a construction site were used to test the proposed method, which
showed success in component extraction and removal of redundant surfaces.
The performance of different non-contact sensors (e.g., 3D laser scanners, photogram-
metry, and 2D cameras) and algorithms (e.g., feature extraction algorithms, Alpha-shape
algorithm, and local entropy-based thresholding algorithm) was reviewed by [9] for the
quality assessment (i.e., dimensional, surface, deflection, and deformation quality assess-
ments) of buildings and civil structures. To enhance the accuracy and applicability of
quality assessments, data fusion-based approaches (e.g., laser scanning data with vision
data, 2D images with depth data, and image data with GPR sensors), as well as more robust
and generic techniques, were suggested. Novel methodologies were developed by [141]
for both the object and damage detection of common structural members from 3D point
cloud data. New skeleton and graph-based object detection approaches were applied to
the segmented point clusters, which involved taking the connectivity information to find
the surfaces apart from one another of the same object. By comparing the fitted objects
from test specimens and test-bed bridges with the acquired point clouds, defects were
located and quantified. An algorithm presented by [142] used TLS for measuring struc-
tural deflection and damage. The TLS data was acquired for the loading and unloading
scenarios, and the plane was fitted for the point cloud using a robust genetic algorithm.
The scanner coordinates were transformed into structural coordinates to be curve fitted for
the loading case. The deflection was estimated and compared between LVDTs. The strain
and deformation of a steel plate under lateral pressure using LiDAR data were evaluated
by [143]. Through a specific interpolation procedure, the point cloud was converted into a
3D mesh model, and the displacements from the initial shape were computed, followed
by the strain calculations. Table 9 summarizes the devices used, type of assessment, and
post-processing method for the laser-based inspection of structural elements.

Table 9. A comprehensive summary of laser-based assessment techniques for other structural elements.

Reference Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


Cyra Cyrax 2500 and Riegl
[121] Deflection assessment Least-square analysis
LMS-Z210
[122] N/A Deflection assessment Least-square analysis
[123] Leica ScanStation 2 Deformation analysis Cross-section comparison
Transformation of
[124] Optech ILRIS-3D Deformation analysis
coordinates differences
[126] FARO Focus 3D Stress analysis Polynomial surface fitting
[127] FARO Focus 3D Deformation analysis Polynomial surface fitting
[128] FARO Photon 120/20 Crack detection and characterization Alpha-Shapes Analysis
[129] N/A Deformation analysis Polynomial surface fitting
[130] N/A Deflection assessment Statistical sampling technique
Clustering algorithms (k-means,
[131] Leica ScanStation C10 Damage detection fuzzy, c-means,
density-based, subtractive)
Sensors 2022, 22, 4610 24 of 32

Table 9. Cont.

Reference Laser-Based Scanning Device Type of Assessment Post-Processing Method(s)


[132] N/A Damage detection Differential analysis
[133] Hexagon Absolute Arm 7325SI Welding inspection Photogrammetric analysis
[134] Leica ScanStation 2 Deformation analysis Polynomial surface fitting
Region growing segmentation and
[135] Leica P20 Structural component identification
the RANSAC
[136] FARO Focus 3D Damage detection Differential analysis
[137] N/A Damage detection Finite element method
[138] Leica AT402 Deformation analysis SfM
[139] N/A Structural component identification Rbnn algorithm
Comparison of as-built and
[140] Leica HDS6100 Structural component identification
planned bim
Skeleton and graph-based
[141] N/A Damage detection
object detection
[142] Leica ScanStation C5 Deflection assessment Genetic algorithm and curve fitting
[143] Leica ScanStation 2 Deflection assessment Interpolation analysis

4. Existing Challenges of LiDAR in SHM


Although LiDAR-based structural assessment has received increased attention in the
SHM field over recent decades, there are still limitations related to the non-contact sensing
process. The most challenging step in LiDAR-based monitoring is acquiring a point cloud
that is high quality, dense, and uniform. Errors in the point cloud can yield redundant data
or gaps in the point cloud, thus also affecting the application and accuracy of algorithms
for detecting structural damage. The quality of data that the MLS or TLS device acquires
is highly dependent on several parameters, including the following: (1) the presence of
occlusions or redundant objects, (2) surface properties, (3) scanner positioning, (4) scanner
specifications, and (5) the environmental conditions.
The first major challenge to acquiring a point cloud that is ready for structural analysis
is the large amount of redundant data that must be filtered. Dependent on the application,
this may include vegetation, people, and non-structural components of the structure or
surface being studied. The presence of this redundant data is problematic, due to the
direct correlation between computational times with the larger datasets. Also, redundant
data can act as obstacles in the scans that prevent the acquisition of data of the objects
of interest. One of the main disadvantages of their scanner-based bridge evaluation was
the presence of occlusions that prevented the entire scan of the target structure in [81].
Limited surficial scans were also conducted by [109], which employed laser scanners for
post-earthquake damage evaluation, but only a few scans were taken of exposed and visible
structural members.
Surface properties also have a significant impact on the quality of the LiDAR scans.
The surface smoothness and general features of the surface may enhance noise within the
point cloud, making it difficult to differentiate between surface deviations that may or
may not impact the structural integrity of the element. A method for identifying cracks
in concrete bridges was developed by [37] but found that there was limited capability
of the proposed method when there were stains that hid the cracks and interrupted the
crack boundaries. Post-processing techniques, as indicated by [41], may be implemented
to reduce the negative effects of surface-based contamination for LiDAR-based crack
quantification. However, the evaluation of the structural health of heritage structures
continues to prove challenging, as concluded by [77], as the larger amount of surface flaws
Sensors 2022, 22, 4610 25 of 32

acquired over time makes it even more difficult for damage and change detection using
LiDAR to provide accurate results.
The quality and completeness of point cloud data are also heavily correlated to the
relative positioning of the scanning device. TLS point clouds can be particularly difficult
to obtain due to overlapping or targetless scans if the field of view or incident angle
of the sensor is not positioned strategically. The proposed method was not suitable for
curvature-based analysis if the noise exceeded a certain threshold in [27]. Similarly, ref. [31]
observed that a larger scan angle increased the error of selecting the same interest point
on surfaces from different scan images, thus also increasing the error for target dimension
measurement. Moreover, the distance from the intended targets also impacts the accuracy
of damage detection methods implementing raw point cloud data. For damage or change
detection, ALS is often subject to larger scanner distances and unfavorable scanning angles.
Airborne sensors were used for post-disaster assessment but found that the scanner could
only acquire information on the roofs and lateral walls of buildings from its position above
in [6]. An unmanned aircraft system was implemented for the structural assessment of
a bridge and used a planned flight path to ensure that the positioning and orientation
uncertainties did not affect the data characteristics in [144].
Scanner specifications, such as the device’s range, the mobility of the device, and
the general settings, are important limitations to be considered for LiDAR-based SHM.
For surface damage detection, it is a common issue that the scanning precision of the
LiDAR device does not allow the identification of smaller defects in the range of several
millimeters [27,81]. When comparing the scanning specifications of TLS and ALS, there are
limitations to using each type of scanner based on the scanner’s requirements. ALS requires
the integration of a GPS, thus rendering GPS-denied areas an issue for acquiring point cloud
data, as described by [145]. However, ref. [40] found that using TLS for change detection
had some disadvantages as well, such as the sensitivity to scanner positioning, fluctuations
in point cloud density, and complex data processing. Additionally, ref. [43] employed both
TLS and MLS in their research and underlined the high noise levels, difficulty in key point
matching for narrow features, and a longer 3D reconstruction process when using MLS.
Finally, the environment in which the LiDAR scans are taken can create obstacles to
performing structural analysis on a high-quality point cloud. The primary environmental
factors include surface illumination (e.g., sunlight and shadows) and weather conditions
(e.g., temperature, humidity, wind, rain, and visibility). Shadows projected by buildings
or vegetation and sunlight contamination all contribute to increased noise within a point
cloud. LiDAR devices work by bouncing laser beams off surrounding objects, therefore,
rain, snow, fog, or dust do not permit scans with high resolution. ALS is often subject to
poorer scanning angles, non-uniform point clouds, and high exposure to the environment;
therefore, airborne LiDAR data can have reduced quality. While using mobile LiDAR, poor
surface illumination or windy environments both highly affected the quality of the point
cloud data and the ability of the proposed algorithms to detect cracks in [145].

5. Future Research Directions of LiDAR-Based SHM Technology


From the literature review, research being conducted using MLS, TLS, and other
LiDAR-based devices over the past two decades has been extensive, resulting in many
advancements to point cloud data capturing and assessment. However, there are still
limitations for point cloud acquisition, data processing, and feature extraction that current
methodologies have not yet addressed. During the data acquisition phase, noise from
environmental factors and low point cloud resolution, due to obscured visual angles,
are still prevalent within this research domain. As a result, data acquisition is often
inaccurate or constrained by the physical parameters of the site under investigation. Future
research should focus on developing optimization methods for parameter selection to
automatically reduce noise and occlusion during data acquisition, depending on the type of
structure. Additionally, best practices for choosing scanning locations and distances should
Sensors 2022, 22, 4610 26 of 32

be developed to ensure the completeness of the scanned data and improve the efficiency of
the scanning process.
Though extracting 3D point cloud from structures is relatively accessible using LiDAR
devices, the complexity and size of the extracted data make the post-collection processing
a complicated and time-extensive endeavor. Most datasets can be reduced based on the
localization of the analysis that is to be conducted; however, current processes require
the manual manipulation of the point cloud to extract relevant information to construct
a 3D model. Methods for the automatic conversion of raw point cloud data into a 3D
model should be explored to maximize the efficiency of post-collection data processing.
Furthermore, the robustness of these data processing techniques must be explored to
ensure accurate 3D models are extracted from the existing point cloud data. Domain-
specific knowledge extracted from objects scanned by LiDAR can be incorporated in the
data post-processing stage to enhance the accuracy of the 3D point cloud conversion.
Therefore, future methods should focus on robust methods that include geometric and
physical parameters of structures for point cloud processing.
Following the processing of point cloud data, the 3D models are used to extract
quantitative and descriptive data about the damaged structure through the use of statistical
models and Artificial Intelligence (AI) techniques. The quality of information extracted
from the model is directly correlated to (1) the quality of the data implemented and (2)
the robustness of the analytical technique used for the investigation. Though many AI
techniques and statistical models have been applied for the evaluation of point cloud-based
models, domain adaptability remain a prevalent issue. Domain adaptation is the ability of
an algorithm that is trained on a source domain (i.e., cracked concrete walls) to perform
on a related target domain (cracked concrete columns). Therefore, future research should
improve the domain adaptation of existing classification techniques to improve robustness
for damage detection for complex structures with a variety of damage types. Furthermore,
the accuracy of the quantification of physical parameters of damages, such as cracks, should
be improved for cases where the damage features are small (<1 cm).

6. Conclusions
The increased prevalence of natural disasters in conjunction with global infrastructure
reaching the end of its service life has spurred the SHM community to develop efficient
monitoring and inspection techniques for structures. The diverse and unique challenges
presented by various structural components, such as roads, bridges, retaining walls, build-
ings, and various other structural elements, have instigated an expanded research effort
into sensor technologies for SHM. Non-contact sensing techniques, particularly LiDAR-
based analysis, have received growing attention with the SHM community recently for
their damage detection abilities through the generation of 3D point clouds. This paper
provides a comprehensive review of LiDAR-based SHM techniques and the analysis of
structural damages using laser-based point cloud data. The development of algorithms for
bridges, tunnels and arch structures, post-disaster reconnaissance, historical and heritage
structures, masonry surfaces, roofs, pavement and roads, structural elements, and walls
have been summarized, and the existing limitations have been discussed. For LiDAR-based
SHM, future research should focus on the development of more computationally efficient,
autonomous models that are connected to the physical domain they are derived from.

Author Contributions: Conceptualization, E.K., K.D. and A.S.; Literature curation, E.K. and K.D.;
writing—original draft preparation, E.K. and A.S.; writing—review and editing, K.D. and A.S.;
visualization, E.K. and K.D.; supervision, A.S.; funding acquisition, E.K. and A.S. All authors have
read and agreed to the published version of the manuscript.
Funding: The authors would like to thank the Undergraduate Student Research Award (USRA)
program of NSERC for providing the financial support to conduct this research through Western
University. The authors also thank the Ontario Ministry of Colleges and Universities for providing
the research funding through the Early Researcher Award to the corresponding author.
Sensors 2022, 22, 4610 27 of 32

Institutional Review Board Statement: Not applicable.


Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Sony, S.; Laventure, S.; Sadhu, A. A literature review of next-generation smart sensing technology in structural health moniotirng.
Struct. Control Health Monit. 2019, 26, e2321. [CrossRef]
2. Mutlib, N.K.; Baharom, S.B.; El-Sharfie, A.; Nuawi, M.Z. Ultrasonic health monitoring in structural engineering: Buildings and
bridges. Struct. Control Health Monit. 2015, 23, 409–422. [CrossRef]
3. Bolourian, N.; Hammad, A. LiDAR-equipped UAV path planning considering potential location of defects for bridge inspection.
Autom. Constr. 2020, 117, 103250. [CrossRef]
4. Pereira, A.; Cabaleiro, M.; Conde, B.; Sanchez-Rodriquez, A. Automatic Identification and Geometrical Modeling of Steel Rivets
of Historical Structure from Lidar Data. Remote Sens. 2021, 13, 2108. [CrossRef]
5. Trias, A.; Yu, Y.; Gong, J.; Moon, F.L. Supporting quantitative structural assessment of highway bridges through the use of LiDAR
scanning. Struct. Infrastruct. Eng. 2021, 18, 824–835. [CrossRef]
6. Dell’Acqua, F.; Gamba, P. Remote Sensing and Earthquake Damage Assessment: Experiences, Limits, and Perspectives. Proc.
IEEE 2012, 10, 100. [CrossRef]
7. Dong, L.; Shan, J. A comprehensive review of earthquake-induced building damage detection with remote sensing techniques.
ISPRS J. Photogramm. Remote Sens. 2013, 84, 85–99. [CrossRef]
8. Coenen, T.B.J.; Colroo, A. A review on automated pavement distress detection methods. Cogent Eng. 2017, 4, 1374822. [CrossRef]
9. Kim, M.-K.; Wang, Q.; Li, H. Non-contact sensing based geometric quality assessment of buildings and civil structures: A review.
Autom. Constr. 2019, 100, 163–179. [CrossRef]
10. Liu, Y.; Hajj, M.; Bao, Y. Review of robot-based damage assessment for offshore wind turbines. Renew. Sustain. Energy Rev. 2022,
158, 112187. [CrossRef]
11. Ranyal, E.; Sadhu, A.; Jain, K. Road condition monitoring using smart sensing and artificial intelligence: A review. Sensors 2022,
22, 3044. [CrossRef] [PubMed]
12. Sony, S.; Dunphy, K.; Sadhu, A.; Capretz, M. A systematic review of convolutional neural network-based structural condition
assessment techniques. Eng. Struct. 2021, 226, 111347. [CrossRef]
13. Mukupa, W.; Roberts, G.W.; Hancock, C.M.; Al-Manasir, K. A review of the use of terrestrial laser scanning application for change
detection and deformation monitoring of structures. Surv. Rev. 2016, 49, 99–116. [CrossRef]
14. Chen, S.-E.; Liu, W.; Bian, H.; Smith, B. 3D LiDAR Scans for Bridge Damage Evaluation. Forensic Eng. 2013, 487–495.
15. Rashidi, M.; Mohammadi, M.; Kivi, S.S.; Abdolvand, M.M.; Truong-Hong, L.; Samali, B. A Decade of Modern Bridge Monitoring
Using Terrestrial Laser Scanning: Review and Future Directions. Remote Sens. 2020, 12, 3796. [CrossRef]
16. Suchocki, C.; Katzer, J. Terrestrial laser scanning harnessed for moisture detection in building materials—Problems and limitations.
Autom. Constr. 2018, 94, 127–134. [CrossRef]
17. Dong, P.; Chen, Q. LiDAR Remote Sensing and Applications; CRC Press: Boca Raton, FL, USA; Taylor and Francis Group: Abingdon,
UK, 2018.
18. McManamon, P. LiDAR Technologies and Systems; SPIE Press: Washington, DC, USA, 2019.
19. Sohn, H.; Park, B. Laser-based structural health monitoring. Encycl. Earthq. Eng. 2015, 1273–1286.
20. Kwiatkowski, J.; Anigacz, W.; Beben, D. A Case Study on the Noncontact Inventory of the Oldest European Cast-iron Bridge
Using Terrestrial Laser Scanning and the Photogrammetric Techniques. Remote Sens. 2020, 12, 2745. [CrossRef]
21. Gonzalez-Jorge, H.; Martinez-Sanchez, J.; Bueno, M.; Arias, P. Unmanned Aerial Systems for Civil Applications: A Review. Drones
2017, 1, 2. [CrossRef]
22. Barbarella, M.; Rosaria De Blasiis, M.; Fiani, M. Terrestrial laser scanner for the analysis of airport pavement geometry. Int. J.
Pavement Eng. 2017, 20, 466–480. [CrossRef]
23. Barbarella, M.; D’Amico, F.; De Blasiis, M.R.; Di Benedetto, A.; Fiani, M. Use of Terrestrial Laser Scanner for Rigid Airport
Pavement Management. Sensors 2017, 18, 44. [CrossRef] [PubMed]
24. Yang, S.; Zhang, Y.; Kaya, O.; Ceylan, H.; Kim, S. Investigation of Longitudinal Cracking in Widened Concrete Pavements. Balt. J.
Road Bridge Eng. 2020, 15, 211–231. [CrossRef]
25. De Blasiis, M.R.; Di Benedetto, A.; Fiani, M. Mobile Laser Scanning Data for the Evaluation of Pavement Surface Distress. Remote
Sens. 2020, 12, 942. [CrossRef]
26. Beshr, A.A.A.; Heneash, O.G.; El-Din Fawzy, H.; El-Banna, M.M. Condition assessment of rigid pavement using terrestrial laser
scanner observations. Int. J. Pavement Eng. 2021, 1–12. [CrossRef]
27. Teza, G.; Galgaro, A.; Moro, F. Contactless recognition of concrete surface damage from laser scanning and curvature computation.
NDTE Int. 2009, 42, 240–249. [CrossRef]
28. Liu, W.; Chen, S.; Hauser, E. LiDAR-Based Bridge Structure Defect Detection. Exp. Tech. 2011, 35, 27–34. [CrossRef]
Sensors 2022, 22, 4610 28 of 32

29. Liu, W.; Chen, S.; Hasuer, E. Bridge Clearance Evaluation Based on terrestrial LIDAR Scan. J. Perform. Constr. Facil. 2012, 26,
469–477. [CrossRef]
30. Watson, C.; Chen, S.-E.; Bian, H.; Hauser, E. Three-Dimensional Terrestrial LIDAR for Operation Bridge Clearance Measurements.
J. Perform. Constr. Facil. 2012, 26, 803–811. [CrossRef]
31. Liu, W.; Chen, S. Reliability analysis of bridge evaluations based on 3D Light Detection and Ranging data. Struct. Control Health
Monit. 2013, 20, 1397–1409. [CrossRef]
32. Riveiro, B.; Gonzalez-Jorge, H.; Varela, M.; Jauregui, D.V. Validation of terrestrial laser scanning and photogrammetry techniques
for the measurement of vertical underclearance and beam geometry in structural inspection of bridges. Measurement 2013, 46,
784–794. [CrossRef]
33. Dai, K.; Boyajian, D.; Liu, W.; Chen, S.E.; Scott, J.; Schmieder, M. Laser-Based Field Measurement for a Bridge Finite-Element
Model Validation. J. Perform. Constr. Facil. 2014, 28, 4014024. [CrossRef]
34. Riveiro, B.; DeJong, M.J.; Conde, B. Automated processing of large point clouds for structural health monitoring of masonry arch
bridges. Autom. Constr. 2016, 72, 258–268. [CrossRef]
35. Stavroulaki, M.E.; Riveiro, B.; Drosopoulos, G.A.; Solla, M.; Koutsianitis, P.; Stavroulakis, G.E. Modelling and strength evaluation
of masonry bridges using terrestrial photogrammetry and finite elements. Adv. Eng. Softw. 2016, 101, 136–148. [CrossRef]
36. Lohmus, H.; Ellmann, A.; Mardla, S.; Idnurm, S. Terrestrial laser scanning for the monitoring of bridge load tests—Two case
studies. Surv. Rev. 2017, 50, 270–284. [CrossRef]
37. Valenca, J.; Puente, I.; Julio, E.; Gonzalez-Jorge, H.; Arias-Sanchez, P. Assessment of cracks on concrete bridges using image
processing supported by laser scanning survey. Constr. Build. Mater. 2017, 146, 668–678. [CrossRef]
38. Sanchez-Rodriguez, A.; Riveiro, B.; Conde, B.; Soilan, M. Detection of structural faults in piers of masonry arch bridges through
automated processing of laser scanning data. Struct. Control Health Monit. 2018, 25, e2126. [CrossRef]
39. Pedro Cortes Perez, J.; Juan de Sanjose Blasco, J.J.D.; Atkinson, A.; Mariano de Rio Perez, L. Assessment of the Structural Integrity
of the Roman Bridge of Alcantara (Spain) Using TLS and GPR. Remote Sens. 2018, 10, 387. [CrossRef]
40. Ziolkowski, P.; Szulwic, J.; Miskiewicz, M. Deformation Analysis of a Composite Bridge during Proof Loading Using Point Cloud
Processing. Sensors 2018, 18, 4332. [CrossRef]
41. Kim, I.-H.; Jeon, H.; Baek, S.-C.; Hong, W.-H.; Jung, H.-J. Application of Crack Identification Techniques for an Aging Concrete
Bridge Inspection Using an Unmanned Aerial Vehicle. Sensors 2018, 18, 1881. [CrossRef]
42. Cha, G.; Park, S.; Oh, T. A Terrestrial LiDAR-Based Detection of Shape Deformation for Maintenance of Bridge Structures.
J. Constr. Eng. Manag. 2019, 145, 4019075. [CrossRef]
43. Chen, S.; Laefer, D.F.; Mangina, E.; Iman Zolanvari, S.M.; Byrne, J. UAV Bridge Inspection through Evaluated 3D Reconstructions.
J. Bridge Eng. 2019, 24, 5019001. [CrossRef]
44. Lee, J.; Lee, K.-C.; Lee, S.; Lee, Y.-J.; Sim, S.-H. Long-term displacement measurement of bridges using a LiDAR system. Struct.
Control Health Monit. 2019, 26, e2428. [CrossRef]
45. Liu, X.; Wang, P.; Lu, Z.; Gao, K.; Wang, H.; Jiao, C.; Zhang, X. Damage Detection and Analysis of Urban Bridges Using Terrestrial
Laser Scanning (TLS), Ground-Based Microwave Interferometry, and Permanent Scatterer Interferometry Synthetic Aperture
Radar (PS-InSAR). Remote Sens. 2019, 11, 580. [CrossRef]
46. Erdelyi, J.; Kopacik, A.; Kyrinovic, P. Spatial Data Analysis for Deformation Monitoring of Bridge Structures. Appl. Sci. 2020,
10, 8731. [CrossRef]
47. Cha, G.; Sim, S.-H.; Park, S.; Oh, T. LiDAR-Based Bridge Displacement Estimation Using 3D Spatial Optimization. Sensors 2020,
20, 7117. [CrossRef] [PubMed]
48. Kermarrec, G.; Kargoll, B.; Alkhatib, H. Deformation Analysis Using B-Spline Surface with Correlated Terrestrial Laser Scanner
observations—A Bridge Under Load. Remote Sens. 2020, 12, 829. [CrossRef]
49. Han, J.-Y.; Guo, J.; Jiang, Y.-S. Monitoring tunnel profile by means of multi-epoch dispersed 3-D LiDAR point clouds. Tunn.
Undergr. Space Technol. 2013, 33, 186–192. [CrossRef]
50. Cheng, Y.-J.; Qiu, W.; Lei, J. Automatic Extraction of Tunnel Lining Cross-Sections from Terrestrial Laser Scanning Point Clouds.
Sensors 2016, 16, 1648. [CrossRef]
51. Zhou, Y.; Wang, S.; Mei, X.; Yin, W.; Lin, C.; Hu, Q.; Mao, Q. Railway Tunnel Clearance Inspection Method Based on 3D Point
Cloud from Mobile Laser Scanning. Sensors 2017, 17, 2055. [CrossRef]
52. Du, L.; Zhong, R.; Sun, H.; Wu, Q. Automatic Monitoring of Tunnel Deformation Based on High Density Point Clouds Data. Int.
Arch. Photogramm. 2017, 42. [CrossRef]
53. Yang, Q.; Zhang, Z.; Liu, X.; Ma, S. Development of Laser Scanner for Full Cross-Sectional Deformation Monitoring of Under-
ground Gateroads. Sensors 2017, 17, 1311. [CrossRef] [PubMed]
54. Xu, X.; Yang, H.; Neumann, I. A feature extraction method for deformation analysis of large-scale composite structures based on
TLS measurement. Compos. Struct. 2018, 184, 591–596. [CrossRef]
55. Xu, T.; Xu, L.; Li, X.; Yao, J. Detection of Water Leakage in Underground Tunnels Using Corrected Intensity Data and 3D Point
Cloud of Terrestrial Laser Scanning. IEEE Access 2018, 6, 2169–3536. [CrossRef]
56. Xu, X.; Yang, H. Intelligent crack extraction and analysis for tunnel structures with terrestrial laser scanning measurement. Adv.
Mech. Eng. 2019, 11, 1687814019872650. [CrossRef]
Sensors 2022, 22, 4610 29 of 32

57. Yang, H.; Xu, X. Intelligent crack extraction based on terrestrial laser scanning measurement. Meas. Control 2020, 53, 416–426.
[CrossRef]
58. Yang, H.; Omidalizarandi, M.; Xu, X.; Neumann, I. Terrestrial laser scanning technology for deformation monitoring and surface
modeling of arch structures. Compos. Struct. 2017, 169, 173–179. [CrossRef]
59. Xu, X.; Yang, H. Network method for deformation analysis of three-dimensional point cloud with terrestrial laser scanning sensor.
Int. J. Distrib. Sens. Netw. 2018, 14, 1550147718814139. [CrossRef]
60. Xu, X.; Yang, H.; Zhang, Y.; Neumann, I. Intelligent 3D data extraction method for deformation analysis of composite structures.
Compos. Struct. 2018, 203, 254–258. [CrossRef]
61. Zhao, X.; Kargoll, B.; Omidalizarandi, M.; Xu, X.; Alkhatib, H. Model Selection for Parametric Surfaces Approximating 3D Point
Clouds for Deformation Analysis. Remote Sens. 2018, 10, 634. [CrossRef]
62. Zhang, W.; Qiu, W.; Song, D.; Xie, B. Automatic Tunnel Steel Arches Extraction Algorithm Based on 3D LiDAR Point Cloud.
Sensors 2019, 19, 3972. [CrossRef]
63. Yang, H.; Xu, X.; Neumann, I. An automatic finite element modelling for deformation analysis of composite structures. Compos.
Struct. 2019, 212, 434–438. [CrossRef]
64. Armesto-Gonzalez, J.; Riveiro-Rodriguez, B.; Gonzalez-Aguilera, D.; Rivas-Brea, M.T. Terrestrial laser scanning intensity data
applied to damage detection for historical buildings. J. Archaeol. Sci. 2010, 37, 3037–3047. [CrossRef]
65. Teza, G.; Pesci, A. Geometric characterization of a cylinder-shaped structure from laser scanner data: Development of an analysis
tool and its use on a leaning bell tower. J. Cult. Herit. 2013, 14, 411–423. [CrossRef]
66. Costanzo, A.; Minasi, M.; Casula, G.; Musacchio, M.; Buongiorno, M.F. Combined Use of Terrestrial Laser Scanning and IR
Thermography Applied to a Historical Building. Sensors 2015, 15, 194–213. [CrossRef] [PubMed]
67. Selbesoglu, M.O.; Bakirman, T.; Gokbayrak, O. Deformation Measurement Using Terrestrial Laser Scanner for Cultural Heritage.
Int. Arch. Photogramm. 2016, 42.
68. Hu, Q.; Wang, S.; Fu, C.; Ai, M.; Yu, D.; Wang, W. Fine Surveying and 3D Modeling Approach for Wooden Ancient Architecture
via Multiple Laser Scanner Integration. Remote Sens. 2016, 8, 270. [CrossRef]
69. Martin Lerones, P.; Olmedo Velez, D.; Gayubo Rojo, F.; Gomez-Garcia-Bermejo, J.; Zalama Casanova, E. Moisture detection in
heritage buildings by 3D laser scanning. Int. Inst. Conserv. Hist. Artist. Work. 2016, 61, 46–54.
70. Korumaz, M.; Betti, M.; Conti, A.; Tucci, G.; Bartoli, G.; Bonora, V.; Gulec Korumaz, A.; Fiorini, L. An integrated Terrestrial Laser
Scanner (TLS), Deviation Analysis (DA) and Finite Element (FE) approach for health assessment of historic structures. A minaret
case study. Eng. Struct. 2017, 153, 224–238. [CrossRef]
71. Castellazzi, G.; D’Altri, A.M.; de Miranda, S.; Ubertini, F. An innovative numerical modeling strategy for the structural analysis
of historical monumental buildings. Eng. Struct. 2017, 132, 229–248. [CrossRef]
72. Rodriguez-Gonzalvez, P.; Fernandez-Palacios, B.J.; Munoz-Nieto, A.L.; Arias-Sanchez, P.; Gonzalez-Aguilera, D. Mobile LiDAR
System: New Possibilities for the Documentation and Dissemination of Large Cultural Heritage Sites. Remote Sens. 2017, 9, 189.
[CrossRef]
73. Jaafar, H.A.; Meng, X.; Sowter, A.; Bryan, P. New approach for monitoring historic and heritage buildings: Using terrestrial laser
scanning and generalised Procrustes analysis. Struct. Control Health Monit. 2017, 27, e1987. [CrossRef]
74. Wilson, L.; Rawlinson, A.; Frost, A.; Hepher, J. 3D digital documentation for disaster management in historic buildings:
Applications following fire damage at the Mackintosh building, The Glasgow School of Art. J. Cult. Herit. 2018, 31, 24–32.
[CrossRef]
75. Campiani, A.; Lingle, A.; Lercari, N. Spatial analysis and heritage conservation; Leveraging 3-D data and GIS for monitoring
earthen architecture. J. Cult. Herit. 2019, 39, 166–176. [CrossRef]
76. Lercari, N. Monitoring earthen archaeological heritage using multi-temporal terrestrial laser scanning and surface change
detection. J. Cult. Herit. 2019, 39, 152–165. [CrossRef]
77. Buill, F.; Amparo Nunez-Andres, M.; Costa-Jover, A.; Moreno, D.; Puche, J.M.; Macias, J.M. Terrestrial Laser Scanner for the
Formal Assessment of a Roman-Medieval Structure—The Cloister of the Cathedral of Tarragona (Spain). Geosciences 2020, 10, 427.
[CrossRef]
78. Pathak, R.; Saini, A.; Wadhwa, A.; Sharma, H.; Sangwan, D. An object detection approach for detecting damages in heritage sites
using 3-D point clouds and 2-D visual data. J. Cult. Herit. 2021, 48, 74–82. [CrossRef]
79. Wood, R.L.; Mohammadi, M.E. Feature-Based Point Cloud-Based Assessment of Heritage Structures for Nondestructive and
Noncontact Surface Damage Detection. Heritage 2021, 4, 775–793. [CrossRef]
80. Tang, P.; Huber, D.; Akinci, B. Characterization of Laser Scanners and Algorithms for Detecting Flatness Defects on Concrete
Surfaces. J. Comput. Civ. Eng. 2011, 25, 31–42. [CrossRef]
81. Mizoguchi, T.; Koda, Y.; Iwaki, I.; Wakabayashi, H.; Kobayashi, Y.; Shirai, K.; Hara, Y.; Lee, H.S. Quantitative scaling evaluation of
concrete structures based on terrestrial laser scanning. Autom. Constr. 2013, 35, 263–274. [CrossRef]
82. Rabah, M.; Elhattab, A.; Fayad, A. Automatic concrete cracks detection and mapping of terrestrial laser scan data. NRIAG J.
Astron. Geophys. 2013, 2, 250–255. [CrossRef]
83. Yang, H.; Xu, X.; Neumann, I. The Benefit of 3D Laser Scanning Technology in the Generation and Calibration of FEM Models for
Health Assessment of Concrete Structures. Sensors 2014, 14, 21889–21904. [CrossRef] [PubMed]
Sensors 2022, 22, 4610 30 of 32

84. Bosche, F.; Guenet, E. Automating surface flatness control using terrestrial laser scanning and building information models.
Autom. Constr. 2014, 44, 212–226. [CrossRef]
85. Kim, M.-K.; Sohn, H.; Chang, C.-C. Localization and Quantification of Concrete Spalling Defects Using Terrestrial Laser Scanning.
J. Comput. Civ. Eng. 2015, 29, 4014086. [CrossRef]
86. Law, D.W.; Holde, L.; Silcock, D. The assessment of crack development in concrete using a terrestrial laser scanner (TLS). Aust. J.
Civ. Eng. 2016, 13, 22–31. [CrossRef]
87. Erkal, B.G.; Hajjar, J.F. Using extracted member properties for laser-based surface damage detection and quantification. Struct.
Control Health Monit. 2020, 27, e2616.
88. Law, D.W.; Silcock, D.; Holden, L. Terrestrial laser scanner assessment of deteriorating concrete structures. Struct. Control Health
Monit. 2018, 25, e2156. [CrossRef]
89. Turkan, Y.; Hong, J.; Laflamme, S.; Puri, N. Adaptive wavelet neutral network for terrestrial laser scanner-based crack detection.
Autom. Constr. 2018, 94, 191–202. [CrossRef]
90. Cho, S.; Park, S.; Cha, G.; Oh, T. Development of Image Processing for Crack Detection on Concrete Structures through Terrestrial
Laser Scanning Associated with the Octree Structure. Appl. Sci. 2018, 8, 2373. [CrossRef]
91. Chen, X.; Li, J.; Huang, S.; Cui, H.; Liu, P.; Sun, Q. An Automatic Concrete Crack-Detection Method Fusing Point Clouds and
Images Based on Improved Otsu’s Algorithm. Sensors 2021, 21, 1581. [CrossRef]
92. Yan, Y.; Mao, Z.; Wu, J.; Padir, T.; Hajjar, J.F. Towards automated detection and quantification of concrete cracks using integrated
images and lidar data from unmanned aerial vehicles. Struct. Control Health Monit. 2021, 28, e2757. [CrossRef]
93. Oskouie, P.; Bcerik-Gerber, B.; Soibelman, L. Automated measurement of highway retaining wall displacements using terrestrial
laser scanners. Autom. Constr. 2016, 65, 86–101. [CrossRef]
94. Riveiro, B.; Lourenco, P.B.; Oliveira, D.V.; Gonzalez-Jorge, H.; Arias, P. Automatic Morphologic Analysis of Quasi-Periodic
Masonry Walls from LiDAR. Comput. Aided Civ. Infrastruct. Eng. 2016, 31, 305–319. [CrossRef]
95. Shen, Y.; Lindenbergh, R.; Wang, J. Change Analysis in Structural Scanning Point Clouds: The Baseline Method. Sensors 2017,
17, 26. [CrossRef] [PubMed]
96. Valero, E.; Bosche, F.; Forster, A. Automatic segmentation of 3D point clouds of rubble masonry walls, and its application to
building surveying, repair and maintenance. Autom. Constr. 2018, 96, 29–39. [CrossRef]
97. Al-Rawabdeh, A.; Aldosari, M.; Bullock, D.; Habib, A. Mobile LiDAR for Scalable Monitoring of Mechanically Stabilized Earth
Walls with Smooth Panels. Appl. Sci. 2020, 10, 4480. [CrossRef]
98. Suchocki, C.; Blaszczak-Bak, W.; Janicka, J.; Dumalski, A. Detection of defects in building walls using modified OptD method for
down-sampling of point clouds. Build. Res. Inf. 2020, 49, 197–215. [CrossRef]
99. Bakula, K.; Pilarska, M.; Salach, A.; Kurczynski, Z. Detection of Levee Damage Based on UAS Data—Optical Imagery and LiDAR
Point Clouds. Int. J. Geo-Inf. 2020, 9, 248. [CrossRef]
100. Kalenjuk, S.; Lienhart, W.; Rebhan, J. Processing of mobile laser scanning data for large-scale deformation monitoring of anchored
retaining structures along highways. Comput. Aided Civ. Infrastruct. Eng. 2021, 36, 678–694. [CrossRef]
101. Khoshelham, K.; Elberink, S.O.; Xu, S. Segment-Based Classification of Damaged Building Roofs in Aerial Laser Scanning Data.
IEEE Geosci. Remote Sens. Lett. 2013, 10, 5. [CrossRef]
102. Awrangjeb, M.; Zhang, C.; Fraser, C.S. Automatic extraction of building roofs using LIDAR data and multispectral imagery.
ISPRS J. Photogramm. Remote Sens. 2013, 83, 1–18. [CrossRef]
103. Kashani, A.G.; Graettinger, A.J. Cluster-Based Roof Covering Damage Detection in Ground-Based Lidar Data. Autom. Constr.
2015, 58, 19–27. [CrossRef]
104. He, M.; Zhu, Q.; Du, Z.; Hu, H.; Ding, Y.; Chen, M. A 3D Shape Descriptor Based on Contour Clusters for Damaged Roof
Detection Using Airborne LiDAR Point Clouds. Remote Sens. 2016, 8, 189. [CrossRef]
105. Kashani, A.G.; Graettinger, A.J.; Dao, T. Lidar-Based Methodology to Evaluate Fragility Models for Tornado-Induced Roof
Damage. Nat. Hazards Rev. 2016, 17, 04016006. [CrossRef]
106. Elberink, S.O.; Shoko, M.; Fathi, S.A.; Rutzinger, M. Detection of collapsed buildings by classifying segmented airborne laser
scanner data. Int. Arch. Photogramm. 2011, 38, W12. [CrossRef]
107. Pesci, A.; Teza, G.; Bonali, E.; Casula, G.; Boschi, E. A laser scanning-based method for fast estimation of seismic-induced building
deformations. ISPRS J. Photogramm. Remote Sens. 2013, 79, 185–198. [CrossRef]
108. Gong, J.; Maher, A. Use of Mobile Lidar Data to assess Hurricane Damage and Visualize Community Vulnerability. Transp. Res.
Rec. 2014, 2459, 119–126. [CrossRef]
109. Mosalam, K.M.; Takhirov, S.M.; Park, S. Applications of laser scanning to structures in laboratory tests and field surveys. Struct.
Control Health Monit. 2014, 21, 115–134. [CrossRef]
110. Vetrivel, A.; Gerke, M.; Kerle, N.; Vosselman, G. Identification of damage in buildings based on gaps in 3D point clouds from very
high resolution oblique airborne images. ISPRS J. Photogramm. Remote Sens. 2015, 105, 61–78. [CrossRef]
111. Zeibak-Shini, R.; Sacks, R.; Ma, L.; Filin, S. Towards generation of as-damaged BIM models using laser-scanning and as-built BIM:
First estimate of as-damaged locations of reinforced concrete frame members in masonry infill structures. Adv. Eng. Inform. 2016,
30, 312–326. [CrossRef]
112. Zhou, Z.; Gong, J.; Guo, M. Image-Based 3D Reconstruction for Post-hurricane Residential Building Damage Assessment.
J. Comput. Civ. Eng. 2016, 30, 04015015. [CrossRef]
Sensors 2022, 22, 4610 31 of 32

113. Aixia, D.; Zongjin, M.; Shusong, H.; Xiaoqing, W. Building Damage Extraction from Post-earthquake Airborne LiDAR Data. Acta
Geol. Sin. 2016, 90, 1481–1489. [CrossRef]
114. Axel, C.; van Aardt, J. Building damage assessment using airborne lidar. J. Appl. Remote Sens. 2017, 11, 46024. [CrossRef]
115. Dai, K.; Li, A.; Zhang, H.; Chen, S.E.; Pan, Y. Surface damage quantification of postearthquake building passed on terrestrial laser
scan data. Struct. Control Health Monit. 2018, 25, e2210. [CrossRef]
116. Moya, L.; Yamazaki, F.; Liu, W.; Yamada, M. Detection of collapsed buildings from lidar data due to the 2016 Kumamoto
earthquake in Japan. Nat. Hazards Earth Syst. Sci. 2018, 18, 65–78. [CrossRef]
117. Zhou, Z.; Gong, J. Automated Analysis of Mobile LiDAR Data for Component-Level Damage Assessment of Building Structures
during Large Coastal Storm Events. Comput. Aided Civ. Infrastruct. Eng. 2018, 33, 373–392. [CrossRef]
118. Zhou, Z.; Gong, J.; Hu, X. Community-scale multi-level post-hurricane damage assessment of residential buildings using
multi-temporal airborne LiDAR data. Autom. Constr. 2019, 98, 30–45. [CrossRef]
119. Janalipour, M.; Mohammadzadeh, A. A novel and automatic framework for producing building damage map using post-event
LiDAR data. Int. J. Disaster Risk Reduct. 2019, 39, 101238. [CrossRef]
120. Akhlaghi, M.M.; Bose, S.; Ebrahim Mohammadi, M.; Moaveni, B.; Stavridis, A.; Wood, R.L. Post-earthquake damage identification
of and RC school building in Nepal using ambient vibration and point cloud data. Eng. Struct. 2021, 227, 1111413. [CrossRef]
121. Gordon, S.J.; Lichti, D.D. Modeling Terrestrial Laser Scanner Data for Precise Structural Deformation Measurement. J. Surv. Eng.
2007, 133, 72–80. [CrossRef]
122. Park, H.S.; Lee, H.M.; Adeli, H.; Lee, I. A New Approach for Health Monitoring of Structures: Terrestrial Laser Scanning. Comput.
Aided Civ. Infrastruct. Eng. 2007, 22, 19–30. [CrossRef]
123. Olsen, M.J.; Kuester, F.; Chang, B.J.; Hutchinson, T.C. Terrestrial Laser Scanning-based Structural Damage Assessment. J. Comput.
Civ. Eng. 2010, 24, 264–272. [CrossRef]
124. Ceylan, A.; Gumus, M. Determination of Deformations as a Result of Seismic Loadings on Two-Dimensional Reinforced Concrete
Frame via Terrestrial Laser Scanners. Exp. Tech. 2014, 38, 19–25. [CrossRef]
125. Lindenbergh, R.; Pietrzyk, P. Change detection and deformation analysis using static and mobile laser scanning. Appl. Geomat.
2015, 7, 65–74. [CrossRef]
126. Cabaleiro, M.; Riveiro, B.; Arias, P.; Caamano, J.C. Algorithm for the analysis of deformations and stresses due to torsion in a
metal beam from LIDAR data. Struct. Control Health Monit. 2015, 23, 1032–1046. [CrossRef]
127. Cabaleiro, M.; Riveiro, B.; Arias, P.; Caamano, J.C. Algorithm for beam deformation modeling from LiDAR data. Measurement
2015, 76, 20–31. [CrossRef]
128. Cabaleiro, M.; Lindenbergh, R.; Gard, W.F.; Arias, P.; van de Kuilen, J.W.G. Algorithm for automatic detection and analysis of
cracks in timber beams from LiDAR data. Constr. Build. Mater. 2017, 130, 41–53. [CrossRef]
129. Yang, H.; Xu, X.; Xu, W.; Neumann, I. Terrestrial Laser Scanning-Based Deformation Analysis for Arch and Beam Structures.
IEEE Sens. J. 2017, 17, 14. [CrossRef]
130. Jafari, B.; Khaloo, A.; Lattanzi, D. Deformation Tracking in 3D Point Clouds Via Statistical Sampling of Direction Cloud-to-Cloud
Distances. J. Nondestruct. Eval. 2017, 36, 65. [CrossRef]
131. Hou, T.-C.; Liu, J.-W.; Liu, Y.-W. Algorithmic clustering of LiDAR point cloud data for textural damage identifications of structural
elements. Measurement 2017, 108, 77–90. [CrossRef]
132. Madrigal, C.A.; Branch, J.W.; Restrepo, A.; Mery, D. A Method for Automatic Surface Inspection Using a Model-Based 3D
Descriptor. Sensors 2017, 17, 2262. [CrossRef]
133. Rodriguez-Gonzalvez, P.; Rodriguez-Martin, M.; Ramos, L.F.; Gonzalez-Aguilera, D. 3D reconstruction methods and quality
assessment for visual inspection of welds. Autom. Constr. 2017, 79, 49–58. [CrossRef]
134. Cheol Jo, H.; Kim, J.; Lee, K.; Sohn, H.-G.; Mook Lim, Y. Non-contact strain measurement for laterally loaded steel plate using
LiDAR point cloud displacement data. Sens. Actuators A Phys. 2018, 283, 362–374.
135. Hong Phong Nguyen, C.; Choi, Y. Comparison of point cloud data and 3D CAD data for on-site dimensional inspection of
industrial plant piping systems. Autom. Constr. 2018, 91, 44–52. [CrossRef]
136. Wang, Q.; Kim, M.-K.; Sohn, H.; Cheng, J.C.P. Surface flatness and distortion inspection of precast concrete elements using laser
scanning technology. Smart Struct. Syst. 2018, 18, 601–623. [CrossRef]
137. Song, M.; Yousefianmoghadam, S.; Mohammadi, M.-E.; Moaveni, B.; Stavridis, A.; Wood, R. An application of finite element
model updating for damage assessment of a two-story reinforced concrete building and comparison with lidar. Struct. Health
Monit. 2018, 17, 1129–1150. [CrossRef]
138. Mistretta, F.; Sanna, G.; Stochino, F.; Vacca, G. Structure from Motion Point Clouds for Structural Monitoring. Remote Sens. 2019,
11, 1940. [CrossRef]
139. Liu, J.; Li, D.; Feng, L.; Liu, P.; Wu, W. Towards Automatic Segmentation and Recognition of Multiple Precast Concrete Elements
in Outdoor Laser Scan Data. Remote Sens. 2019, 11, 1383. [CrossRef]
140. Maalek, R.; Lichti, D.D.; Ruwanpura, J.Y. Automatic Recognition of Common Structural Elements from Point Clouds for
Automated Progress Monitoring and Dimensional Quality Control in Reinforced Construction. Remote Sens. 2019, 11, 1102.
[CrossRef]
141. Erkal, B.G.; Hajjar, J.F. Laser-based surface damage detection and quantification using predicted surface properties. Autom. Constr.
2017, 83, 285–302. [CrossRef]
Sensors 2022, 22, 4610 32 of 32

142. Maru, M.B.; Lee, D.; Cha, G.; Park, S. Beam Deflection Monitoring Based on a Genetic Algorithm Using Lidar Data. Sensors 2020,
20, 2144. [CrossRef]
143. Jo, H.C.; Sohn, H.-G.; Lim, Y.M. A LiDAR Point Cloud Data-Based Method for Evaluating Strain on a Curved Steel Plate Subjected
to Lateral Pressure. Sensors 2020, 20, 721. [CrossRef] [PubMed]
144. Morgenthal, G.; Hallermann, N.; Kersten, J.; Taraben, J.; Debus, P.; Helmrich, M.; Rodehorst, V. Framework for Automated
UAS-Based Structural Condition Assessment of Bridges. Autom. Constr. 2019, 97, 77–95. [CrossRef]
145. Dorafshan, S.; Maguire, M. Bridge inspection: Human performance, unmanned aerial systems and automation. J. Civ. Struct.
Health Monit. 2018, 8, 443–476. [CrossRef]

You might also like