0% found this document useful (0 votes)
31 views29 pages

Water 15 00129

Uploaded by

nam2024conrong
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
31 views29 pages

Water 15 00129

Uploaded by

nam2024conrong
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 29

water

Article
Flood Vulnerability Study of a Roadway Bridge Subjected
to Hydrodynamic Actions, Local Scour and Wood
Debris Accumulation
Mirko Kosič, Andrej Anžlin and Valentina Bau’ *

Slovenian National Building and Civil Engineering Institute (ZAG), Dimičeva ulica 12, 1000 Ljubljana, Slovenia
* Correspondence: valentina.bau@zag.si

Abstract: The increased occurrence and intensity of flooding events have represented a real threat
to bridge reliability and end-user safety. As flood vulnerability assessment is a valuable tool for
enhancing the resilience of bridges to climate change, it is of interest to push the development of such
methods. To this end, a computationally efficient methodology to assess the flood vulnerability of a
bridge was developed and implemented in a case study. A particular focus was devoted to modelling
wood debris loads on the bridge pier, for which two different approaches were implemented. The
first is a standards-based approach, whereas the second is based on up-to-date research data. The
results indicate that the second approach is less conservative as it leads up to a 40% higher exceedance
probability for the considered limit states. The interaction between wood debris loads and local
scour was also examined and proved to have a relevant impact on the vulnerability of the bridge.
These results highlight the shortcomings of the existing standards in providing accurate results. It is
perceived that not only will the new quantitative tool be valuable in ensuring optimal bridge design,
but it will also be beneficial for assessing bridge risk mitigation measures.

Keywords: flooding; vulnerability; wood debris; local scour; hydrodynamic load; stochastic
approach; bridges
Citation: Kosič, M.; Anžlin, A.; Bau’,
V. Flood Vulnerability Study of a
Roadway Bridge Subjected to
Hydrodynamic Actions, Local Scour
1. Introduction
and Wood Debris Accumulation.
Water 2023, 15, 129. https:// Bridges have a fundamental role in transportation infrastructure and the overall
doi.org/10.3390/w15010129 world economy; as such, their targeted performance must be met. Complying with the
requirements for modern bridge design requires accurate modelling and a careful analysis
Academic Editors:
of the vulnerability to natural hazards. Local scour and other hydraulic-related processes
José González-Cao,
(e.g., channel migration, flood-associated loads, and wood debris accumulation) have been
D. Fernández-Nóvoa and
Orlando Garcia Feal
proven to be the main reasons for bridge collapse [1], thus affecting the functionality of
the transport system. Evidence of such a pivotal danger is the scientific literature, which
Received: 2 December 2022 has reported several failures of infrastructures triggered by flooding events in the USA,
Revised: 20 December 2022 the United Kingdom, and Europe [2–6]. In this last decade, scientists and practitioners
Accepted: 22 December 2022 have put considerable effort into focusing on the vulnerability of riverine bridges to natural
Published: 29 December 2022
hazards, particularly flooding and local scouring events [7].
The structural integrity of bridges that cross rivers can be compromised by hydrody-
namic forces as well as by the loading effect caused by the accumulation of floating wood
Copyright: © 2022 by the authors.
debris at the piers. Most commonly, the type of debris that endangers bridge safety consists
Licensee MDPI, Basel, Switzerland. of waterborne debris (e.g., tree trunks and limbs) that are transported by the water current
This article is an open access article during flooding events [8]. The presence of debris jams at the piers of bridges may cause
distributed under the terms and an increase in the upstream water levels (afflux) and a backwater effect, thus posing an
conditions of the Creative Commons additional flood hazard for the nearby areas [9]. Wood debris, once stranded at the piers,
Attribution (CC BY) license (https:// can also lead to an increase in the hydrodynamic actions and trigger an exacerbation of the
creativecommons.org/licenses/by/ local scour [10]. Local scour is an equally threatening process that can lead to uncovering
4.0/). the piers’ foundations and reducing their lateral stiffness and load-bearing capacity [11].

Water 2023, 15, 129. https://doi.org/10.3390/w15010129 https://www.mdpi.com/journal/water


Water 2023, 15, 129 2 of 29

Local scour at bridges results from the removal of sediment from around the base of the
pier caused by the onset of horseshoe vortices [12]. The action exerted by local scour has
been responsible for 60% of bridge failures in the United States, according to [13]. Despite
a variety of failure modes associated with local scour, vertical and lateral modes were most
frequently observed [14]. Therefore, it is evident that flood-induced hydrodynamic and
debris impact can undermine the structural safety of bridges and lead to severe structural
damage and consequent economic losses [15].
In order to assess the vulnerability of bridges to floods and develop resilient coun-
termeasure strategies, the implementation of fragility curves represents a reliable and
consolidated methodology [16]. Unlike deterministic approaches, vulnerability analysis
consists of computing the conditional probability of exceedance of a defined damage state
for a considered hazard or a combination of them. While the vulnerability of bridges
to seismic performance has been widely investigated [17–21], little is known about how
flooding events and related processes can simultaneously affect the structural vulnerability
of a bridge. Moreover, although the flood-related risk was included in seismic vulnerability
analysis, its contributions were not considered the main cause of bridge failure [22,23].
Nevertheless, due to the incumbency of changes in global climate, recent studies have
started to implement a more accurate vulnerability analysis of bridges against floods. For
instance, the authors of [24] carried out a flood vulnerability analysis by taking into account
the simultaneous contribution of bridge scour, structural deterioration of the steel rein-
forcements and piles, and the increased water pressure due to debris accumulation around
bridge piers. Another approach was presented by [25], who included the geotechnical
uncertainties provided by the foundation design in the vulnerability analysis. Additionally,
the authors of [26] presented a holistic approach where the fragility curves were imple-
mented examining different combinations of scouring conditions and earthquake loadings
for a reinforced concrete bridge with shallow foundations. Another efficient framework
has been introduced by [27], who identified the possible modes to bridge failure caused by
scour-related processes. An interesting perspective is also offered by [28], who provides a
performance-based engineering framework on a flood fragility analysis of a bridge sub-
jected to various loading scenarios and local scour. Fragility curves were also employed
to illustrate the exceedance probability of the limit states of a masonry bridge exposed
to scour-induced support rotation [29]. In spite of the lack of a classic fragility analysis
approach, the authors of [30] offer a comprehensive and innovative analysis of local scour.
Using artificial intelligence techniques, the authors implement a probabilistic methodol-
ogy that can predict scour depths created by regular waves on groups of piles. It is also
discussed that probabilistic approaches should be able to incorporate uncertainties for the
most critical parameters that influence local scour depths (e.g., pile diameter, average flow
velocity) [31].
Driven by these emerging research perspectives, the presented framework refers
to a vulnerability analysis methodology that aims at using a probabilistic approach to
grasp some of the uncertain variables that affect bridge resilience to floods. Particular
focus is placed on the procedure for modelling wood debris accumulation on the bridge
piers. The vulnerability analysis of this study is implemented for both the serviceability
and ultimate limit state to provide an overall view of the structure’s performance. The
variability of material and geotechnical properties, hydrodynamic loading, wood debris size
accumulation, and scour magnitude is taken into account to illustrate different scenarios
through the analytical implementation of vulnerability surfaces and two failure modes. The
process of a random sampling of the variables is performed with a very efficient stratified
sampling approach (i.e., Latin Hypercube Sampling (LHS)), thus abandoning the typical
Monte Carlo simulation.
The strength and stiffness of the construction materials (e.g., concrete and reinforcing
steel) are some of the major sources of uncertainties in civil engineering structures. Their
assigned values are critical determinants of the behavior of a structure [32]. Thorough
studies have been undertaken to assess the statistical parameters and models of materials
Water 2023, 15, 129 3 of 29

properties [33,34], thus providing solid support when considering such uncertainties in
vulnerability analysis. Similarly, the natural variability of soil properties and especially
the influence of soil spatial variability on structural performance have also been widely
investigated [35,36]. The same cannot be said when considering the uncertainties provided
by the actions of flood and wood debris loading and local scouring. The gap in information
and criteria to incorporate flood-related processes into the vulnerability analysis of bridges,
as opposed to earthquakes, may be due to the difficulties related to monitoring such
processes. The complexity of hydraulic data, the high cost of monitoring techniques,
and the stochasticity inherent in flood-related processes may discourage researchers and
practitioners from exploring such crucial conditions for bridge vulnerability evaluation [37].
This work aims to implement the aforementioned methodology and create a reliability
analysis framework to investigate flood vulnerability through a real-bridge case study.
The originality of this paper stands in offering an analysis framework that considers the
addition of both hydraulics and geotechnical model uncertainties to the ones that are
commonly established in the vulnerability analysis of bridges (e.g., material uncertainties).
The novelty of this paper also relies upon the way the vulnerability analysis is addressed.
The largest part of the existing studies on bridge vulnerability analysis has often modelled
the hydrodynamic and wood debris loads by referring to government guidelines or bridge
design specifications (e.g., AASHTO LRFD Bridge Design Specification [25]). However,
such references are rather simplistic and differ widely due to a lack of consistency in
cross-countries methodology comparisons [13,38]. Furthermore, the use of regulations
and standard specifications can lead to an inherent conservatism that is not beneficial for
vulnerability analysis since it overestimates the probability of reaching the considered
damage limit state analysis. For this reason, the goal of this study is to utilize the latest
research outputs to provide a different perspective and a source of comparison to the most
relevant existing guidelines.

2. Materials and Methods


2.1. Material Properties and Input Data
2.1.1. Geometry and Materials
The case study is a historic roadway riverine steel bridge located in central Croatia.
The bridge has two main spans, a total length of approximately 97 m and a total width of
12.4 m. It supports two traffic lanes and a pedestrian zone. Information on the geometry
and materials of the bridge was retrieved from relevant as-built construction informa-
tion, a recent retrofitting project, and a drone survey. The overview and the geometrical
characteristics of the bridge are illustrated in Figures 1 and 2, respectively.
The two main girders have riveted I-shape cross-sections with flanges whose thickness
variation follows the bending moment line. The main girders are joined together every
4.05 m by riveted transverse beams that are connected with seven I-shaped stringers that,
in turn, support the reinforced concrete (RC) slab of the deck (Figure 1c).
At the abutments, the bridge is supported by longitudinal roller bearings with a
diameter of 20 cm (Figure 1d). The transverse displacement of each bearing is restrained
with shear keys of size 0.03 × 0.06 m. At the pier, the translation of the deck is restrained
in all directions by the presence of shear pins with a diameter of 0.05 m. The pier and the
abutments are made of reinforced concrete and are covered with 15 cm stone cladding. The
middle pier has a rectangular section with rounded ends (see Figure 2b) and is founded
on two large well foundations of 6.5 m diameter (D). Due to the unavailability of as-built
construction plans, the depth of each foundation is assumed to be equal to that used in
the pre-existing bridge from 1889 (L = 5.8 m), for which partial historic construction plans
were obtained.
Water 2023,
Water 15, x129
2023, 15, FOR PEER REVIEW 4 4ofof 30
29

(a) (b)

(c) (d)
Figure
Figure 1.
1.Overview
Overviewofofthe case-study
the case-studybridge: (a) (a)
bridge: sideside
view, (b) (b)
view, detailed viewview
detailed of the
of middle pier and
the middle pier
well foundations, (c) view below the bridge deck, and (d) detailed view of the roller bearing
and well foundations, (c) view below the bridge deck, and (d) detailed view of the roller bearing at the
at
abutment.
the abutment.

At
Thethe abutments,characteristics
mechanical the bridge is supported by longitudinal
of the superstructure wereroller bearings
gained throughwiththea di-
de-
ameter of 20 cm (Figure 1d). The transverse displacement of each bearing
structive testing performed within the retrofitting project. The steel grade used for the is restrained
with shear keysisofestimated
superstructure size 0.03to × 0.06 m. At thetopier,
be equivalent S235the translation
according of the deck
to EN-1993 [39],iswhereas
restrainedthe
in all directions
grade of concrete byofthe
thepresence of shear apins
deck is assigned gradewith a diameter
of C25/30 of 0.05 to
according m.EN-1992-1
The pier and [40].the
In
abutments
the absenceare of made of reinforced
the material concreteused
characteristics and for
arethe
covered
bridgewith
piers15and
cmfoundations,
stone cladding. the
The
grade middle pier has isa assumed
of the concrete rectangularto besection
the samewithas rounded ends deck
for the bridge (see (C25/30).
Figure 2b) and is
founded on two large well foundations of 6.5 m diameter (D). Due to the unavailability of
2.1.2. Geotechnical
as-built constructionData plans, the depth of each foundation is assumed to be equal to that
used Thein the pre-existing
input data for the bridge from 1889
modelling (L soil–structure
of the = 5.8 m), for which partial(SSI)
interaction historic
at theconstruc-
bridge’s
tion plans
central were obtained.
foundation were obtained from a Cone Penetration Test investigation (CPT) and a
The mechanical
borehole performed near characteristics
the bridge’s of abutment.
the superstructure were gained
The soil profile consiststhrough
of a topthe6 mde-of
structive
clay layertesting
with anperformed
underlyingwithin the retrofitting
medium-dense project.
sand layer. The steel grade
Investigations used directly
conducted for the
near the foundation
superstructure of the central
is estimated pier are not
to be equivalent toavailable. However,
S235 according geophysical
to EN-1993 [39], profiling
whereas
indicated
the grade that the stratigraphy
of concrete of the deckoccurring near athe
is assigned bridge’s
grade abutment
of C25/30 also occurs
according near the
to EN-1992-1
foundation.
[40]. Hence,of
In the absence thethe
foundation is assumed to used
material characteristics be located
for theinbridge
the medium-dense
piers and founda- sand
layer. the
tions, Thegrade
estimated
of thesoil characteristics
concrete is assumed forto
modelling
be the same the as
soil–structure
for the bridge interaction at the
deck (C25/30).
central pier are summarized in Table 1.
Water 2023, 15, 129 5 of 29
Water 2023, 15, x FOR PEER REVIEW 5 of 30

(a)

(b)
Figure
Figure2.
2.(a)
(a)Cross-section
Cross-sectionof
ofthe
thecase-study
case-studybridge;
bridge; (b)
(b) cross-section
cross-section A-A
A-A at
at the
the base
base of
of the
the pier.
pier.

2.1.2. Geotechnical Data


Table 1. Soil characteristics used for modelling the soil–structure interaction at the central pier of
The input data for the modelling of the soil–structure interaction (SSI) at the bridge’s
the bridge.
central foundation were obtained from a Cone Penetration Test investigation (CPT) and a
borehole performed near the bridge’s abutment. The soil profileValue
Parameters consists of a top 6 m of
clay layer with an underlying medium-dense sand layer. Investigations
Effective shear angle ( ϕ0 ) 40◦
conducted di-
rectly near the foundation
Effective cohesionof the
(c0 )central pier are not available. However,
0 kPa geophysical pro-
filing indicated
Soilthat
unitthe stratigraphy
weight (γ ) occurring near the bridge’s abutment
19.8 kN/m 3 also occurs
0
near the foundation.
PoissonHence,
s ratio (the
ν ) foundation is assumed to be located in the medium-dense
0.30
sand layer. TheElastic modulus
estimated soil( Echaracteristics
) 50 Mpa
for modelling the soil–structure interaction
Shear modulus ( G )
at the central pier are summarized in Table 1. 19 Mpa
Equivalent SPT blows ( N60 ) 19

2.1.3. Hydraulic Data


The hydraulic data was retrieved based on existing numerical simulation results. In
particular, the data from the simulations refer to the maximum daily values computed for
extreme past flooding events that occurred in 2014 and 2015, respectively. In Figure 3a,b, the
relationships between the water height (H) and water discharge (Q), and between the water
height (H) and flow velocities (v), respectively, illustrate the complex hydraulic conditions
encountered at the location. The data dispersion suggest that the hydraulic conditions are
The vulnerability assessment is performed within the bounds marked by the 95%
prediction intervals (PI) obtained from the regression of the data with the steady-flow
Manning’s equation [41]. Knowing the intervals within which new observations are likely
to fall makes it possible to explore flooding scenarios that are also more severe than the
ones registered in 2014 and 2015, providing an idea of the limit states that the structure
Water 2023, 15, 129 6 of 29
may encounter if certain water heights and flow velocities were reached.
Note that the computations are carried out considering a limit for the water heights
of 10.8 m, which corresponds to the full submergence of the deck of the bridge, and a
strongly
minimum affected by the variable
flow velocity of 0.1 backwater effect ofbelow
m/s, as velocities the tributary riverare
this value, located downstream
irrelevant for the
from the bridge.
purpose of this study.

(a) (b)

Figure 3. Results obtained from the numerical simulations of two flooding events. (a) Relation
between water height (H) and discharge (Q); (b) relation between water velocity (v) and water height
(H). The steady flow relationship and the 95% prediction interval curves are also represented.

Furthermore, the limited amount of data makes it hard to detect a particular hysteretic
pattern or any particular trend in the data. Given the circumstances, the vulnerability
analysis cannot rely on a unique H − v relationship and is therefore implemented for
unsteady flow conditions (i.e., with consideration of both flow velocities and water heights
as input parameters).
The vulnerability assessment is performed within the bounds marked by the 95%
prediction intervals (PI) obtained from the regression of the data with the steady-flow
Manning’s equation [41]. Knowing the intervals within which new observations are likely
to fall makes it possible to explore flooding scenarios that are also more severe than the
ones registered in 2014 and 2015, providing an idea of the limit states that the structure
may encounter if certain water heights and flow velocities were reached.
Note that the computations are carried out considering a limit for the water heights
of 10.8 m, which corresponds to the full submergence of the deck of the bridge, and a
minimum flow velocity of 0.1 m/s, as velocities below this value, are irrelevant for the
purpose of this study.

2.2. Methodology and Modelling Approach


2.2.1. Overview
The complete flowchart of the proposed framework is presented in Figure 4. As can
be seen in the flowchart, both deterministic and stochastic modelling approaches are used.
The vulnerability analysis begins with determining the external forces acting on the
bridge (e.g., hydrodynamic forces, gravity forces, and loads resulting from wood debris
accumulation) that, along with other data inputs (geotechnical data, bridge geometry, and
local scour assessment), are necessary to implement the structural finite-element modelling
and the soil–structure interaction. These parameters combined represent the deterministic
model of the bridge, which specifies the computation of the structural behavior of the
bridge based on a particular set of input information. Details regarding the structural
finite-element, the soil–structure interaction, load and local scour modelling are presented
in Sections 2.2.2–2.2.5, respectively.
Water 2023,
Water 15, x129
2023, 15, FOR PEER REVIEW 8 7ofof 30
29

Figure 4. Flowchart illustrating the procedure that is implemented for the flood vulnerability anal-
Figure 4. Flowchart illustrating the procedure that is implemented for the flood vulnerability analysis
ysis of the bridge.
of the bridge.

In
The this study, the serviceability
deterministic model is converted limit state (SLS) as well
into stochastic in theas second
the ultimate
step. In limit
thisstate
way,
(ULS)
it becomes possible to include in the model the variability of the input parametersdur-
are analyzed. The first is assumed to be related to the operability of the bridge that
ing flooding,
defines whereas thedistribution
the probabilistic latter is related
of thetoloading
the capacity
actionsof and
the bridge to withstand
resistance flood-
of the bridge. In
ing
thisevents
study, without collapsing.
it is assumed that the probabilistic distribution of the limit states of the bridge
is related to the variability of the material characteristics (concrete strength, steel yield
2.2.2. Structural
strength Finite-Element
and elastic modulus),Modelling
geotechnical properties of the soil (soil shear modulus,
effective
A 3Dshear angle,model
numerical and unit weight),
of the bridge hydrodynamic
is generated in forces, local scour depths,
the OpenSeesPy Libraryand [42],size
an
of wood debris
open-source accumulation.
Python3 interpreterThefordetails regarding
OpenSees [43]. the
Theconsidered
modelling uncertainty parameters
approach relies on lin-
are elastic
ear provided beam in Section
elements 2.2.6.
and zeros length-elements soil springs for modelling the soil–
As part of the next step,
structure interaction. A user-defined the propagation
toolbox isofdeveloped
the uncertainty of the model’s
in the Python inputsin
environment is
implemented through the use of Latin Hypercube Sampling (LHS),
order to carry out a fully parameterized analysis for bridges subjected to hydrodynamic which is described in
Section
and wood 2.2.7.
debrisTheloads
advantage of this method
in combination is that
with the it reduces
scouring actiontheonnumber of simulations
the piers.
significantly, which makes it more computationally efficient
A schematic representation of the model is presented in Figure 5. Each bridgethan Monte Carlo approach.
mem-
ber isAfter the model
modelled with uncertainties
linear elastic are propagated,
Timoshenko it is possibleelements.
beam-column to generate theload
The fragility sur-
transfer
between individual beam elements is achieved by the definition of fictitious infinitely stiffa
faces of the bridge. The fragility surfaces indicate the probability of reaching or exceeding
designated
and weightless limitelements
state (LS)(stiff
for aelements),
given levelwhich
of floodareintensity (Intensity
used to connect Measure,
the centroidIM). Gen-
lines of
erally, the exceedance probability of a designated LS is represented
the elements. A more precise distribution of the forces in the bearing at the abutments andby the so-called fragility
curve.
the However,
middle pier isingained
this study, due to thethe
by positioning lackbearing
of a univocal − v relationship,
at theirHactual the IM
location in plan and is
defined based on both parameters (i.e., H and v).
elevation using rigid elements (see Figure 5). The deck of the bridge is modelled with ana
Therefore, the probability of exceeding
designated LS is expressed as P( LS| H, v) and is represented by a fragility surface.
equivalent beam element. The boundary conditions at the abutments simulate the roller
In this study, the serviceability limit state (SLS) as well as the ultimate limit state (ULS)
bearings’ behavior (i.e., free translation in the longitudinal direction and free rotations).
are analyzed. The first is assumed to be related to the operability of the bridge during
The bearings at the pier are modelled as pinned.
Water 2023, 15, 129 8 of 29

flooding, whereas the latter is related to the capacity of the bridge to withstand flooding
events without collapsing.

2.2.2. Structural Finite-Element Modelling


A 3D numerical model of the bridge is generated in the OpenSeesPy Library [42],
an open-source Python3 interpreter for OpenSees [43]. The modelling approach relies
on linear elastic beam elements and zeros length-elements soil springs for modelling the
soil–structure interaction. A user-defined toolbox is developed in the Python environment
in order to carry out a fully parameterized analysis for bridges subjected to hydrodynamic
and wood debris loads in combination with the scouring action on the piers.
A schematic representation of the model is presented in Figure 5. Each bridge member
is modelled with linear elastic Timoshenko beam-column elements. The load transfer
between individual beam elements is achieved by the definition of fictitious infinitely stiff
and weightless elements (stiff elements), which are used to connect the centroid lines of
the elements. A more precise distribution of the forces in the bearing at the abutments and
the middle pier is gained by positioning the bearing at their actual location in plan and
elevation using rigid elements (see Figure 5). The deck of the bridge is modelled with an
equivalent beam element. The boundary conditions at the abutments simulate the roller
Water 2023, 15, x FOR PEER REVIEW 9 of 30
bearings’ behavior (i.e., free translation in the longitudinal direction and free rotations).
The bearings at the pier are modelled as pinned.

Figure5.5.Scheme
Figure Scheme
of of
thethe numerical
numerical model
model ofcase-study
of the the case-study bridge.
bridge.

2.2.3. Soil–Structure Interaction Modelling


2.2.3. Soil–Structure Interaction Modelling
The soil–structure interaction at the well foundations is modelled with a lumped-
The soil–structure interaction at the well foundations is modelled with a lumped-
spring approach. Due to the small ratio between the depths and their diameters𝐿 5.8
5.8approach. Due to the small ratio between the depths and their diameters ( =
spring
L
( D = 6.5 = 0.90), the wells are modelled as embedded shallow foundations. In such𝐷 6.5 =
a0.90),
case, the
theSSI is modelled
wells with six
are modelled mutually independent
as embedded springs located
shallow foundations. at the
In such base of
a case, the SSI is
each well. Figure 5 shows only a scheme of springs with a unique definition,
modelled with six mutually independent springs located at the base of each well. Figurei.e., two hori-
zontal andonly
5 shows two rotational
a schemesprings have with
of springs the same characteristics
a unique in two
definition, i.e.,different directions.
two horizontal and two
The springs are aggregated into a single zero-length element, which connects
rotational springs have the same characteristics in two different directions. The springs the base
of the well with a fully restrained node. The static stiffness of the foundation springs is
are aggregated into a single zero-length element, which connects the base of the well with
calculated following the solution found for embedded cylindrical foundations by Pais and
a fully restrained node. The static stiffness of the foundation springs is calculated follow-
Kausel [44]. The solution considers soil as a linear-elastic material and the foundation as a
ing the
rigid body,solution
with thefound for embedded
presence cylindrical
of a non-slip foundations
interface between by Pais and
the foundation and Kausel [44]. The
the soil.
solution
The considers
influence soil as a linear-elastic
of the foundation embedmentmaterial and account
is taken into the foundation
through as a rigid body, with
amplification
the presence
factors, of aapplied
which are non-slipto interface between
the solution obtainedthefor
foundation and the soil.
surface foundations. The The influence of
vertical
the foundation embedment is taken into account through amplification factors, which are
applied to the solution obtained for surface foundations. The vertical (𝐾𝑉𝑒 ), horizontal (𝐾𝐻𝑒 ),
rocking (𝐾𝑅𝑒 ), and torsional (𝐾𝑇𝑒 ), stiffness of the (embedded) well foundation are com-
puted as follows:
Water 2023, 15, 129 9 of 29

e ), horizontal (K e ), rocking (K e ), and torsional (K e ), stiffness of the (embedded) well


(KV H R T
foundation are computed as follows:

e 0
  4 G Rf  
KV = KV 1 + 0.54 DE /R f = 1 + 0.54 DE /R f (1)
1−ν
  8 G Rf  
K eH = K0H 1 + DE /R f = 1 + DE /R f (2)
2−ν
8 G Rf 3
    3 
e 0
K R = K R 1 + DE /R f = 1 + 2.3 DE /R f + 0.58 DE /R f (3)
3 (1 − ν )
  16 G R f 3  
KTe = KT0 1 + DE /R f = 1 + 2.67 DE /R f (4)
3
where R f is the radius, DE is the embedment depth of the foundation, and G and ν are the
shear modulus and Poisson’s ratio of the soil, respectively. The superscript e denotes the
stiffness of the embedded foundation, whereas the stiffness of the surface foundation is
denoted by the 0 superscript.

2.2.4. Loads Modelling


Loads on the structure come from gravity (self-weight and permanent loads), hydro-
dynamic forces (drag Fd and lift forces Fl ), buoyancy forces, and wood debris accumulation.
In the numerical model, each load is computed per unit length and is applied uniformly.
The loads are combined to produce the maximum possible overturning action on the
bridge, which occurs when horizontal hydrodynamic and wood debris loads are combined
with an uplift action on the deck. In such a case, the hydrodynamic uplift of the deck is
combined with the buoyancy. For this reason, the simultaneous presence of traffic loads is
not considered, as this would result in a stabilizing action on the bridge.
The hydrodynamic loads are only applied to the areas of the bridge that are not
affected by debris loads to avoid the duplication of hydrodynamic actions.

Hydrodynamic Forces
The hydrodynamic forces are computed according to the Australian standard AS5100.2-
2004 [45]. Due to its comprehensive approach, the Australian standard is chosen over the
ASSHTO standard, since it provides additional guidelines for quantifying the hydrody-
namic loading on the bridge’s deck. The drag force in the direction of the flow (in kN) is
defined as follows:
Fd = 0.5 Cd v2 Ad (5)
where Cd is the drag coefficient, v is the average water velocity and Ad is projected wetted
area in the direction of the flow. According to the Australian standard, for piers with
semi-circular nosing such as the case study bridge’s pier, the coefficient Cd equals 0.7. For
the deck of the bridge, the coefficient Cd is instead computed as a function of the relative
submergence of the deck (Sr ) and proximity ratio (Pr ), which are, respectively, defined as:

dwgs y gs
Sr = and Pr = (6)
dsp dss

where dwgs is the distance from the girder soffit to flood water surface, dsp is the wetted
depth of the deck (including any railing and parapets), y gs is the average vertical distance
from the girder soffit to the riverbed, and dss is the wetted depth of the solid superstructure
(excluding railing but including parapets).
The lift force on the bridge pier and deck (in kN) is computed with the following
equation:
Fl = 0.5 Cl v2 Al (7)
where Cl is the lift coefficient, and Al is the wetted area perpendicular to the flow.
Water 2023, 15, 129 10 of 29

In this study, the lift force is only computed in relation to the deck in the cases when it
is submerged. The lift force on the pier is neglected as the pier is oriented in the direction
of the flow. The coefficient Cl for the bridge deck depends on its relative submergence (Sr ).
As indicated in the literature ([24,25,46,47]), the distribution of the hydrodynamic
forces over the water height can be well approximated by a triangular shape. The triangular
distribution of the forces yields a lower safety margin than a uniform distribution because
it results in a larger overturning moment above the base of the foundation.
The buoyancy force is computed as the weight of the displaced water and is applied
concurrently with the upward lift force on the bridge’s deck. The magnitude of the buoy-
ancy force varies according to the level of submergence of the deck. The buoyancy force
increases with the level of submergence of the deck and reaches its maximum value when
the deck is fully submerged.

Wood Debris Forces


Wood debris load is modelled with two different approaches to benchmark its impact
on the flood vulnerability of the bridge. The first approach is based on the guidelines
provided by the Australian standard (AS5100.2-2004), whereas the second one refers to a
recent model drawn from the scientific literature that is herein introduced as Panici and
Almeida’s model [48]. For the sake of simplicity, henceforth, the word wood debris will be
simply referred to as debris.
As per Australian standard, the debris mat has a rectangular shape, with height and
width indicated. In the absence of more accurate estimates, the minimum height of the
wood debris mat (Hd ) is set to be between 1.2 and 3 m. Depending on whether the debris
accumulation is acting on the bridge deck or the pier, the width of the debris mat changes
(Wd ). In the case of debris acting on the bridge’s piers, the width of the debris mat is
assumed to be the smallest value between one-half of the sum of the adjacent spans’ length
(S L ) and 20 m. In the case of debris acting on the bridge deck, the projected width of the
debris mat is equal to the projected length of the deck (DL ). For the purpose of this study,
the height of the debris is set at 2.1 m, which is the mean value of the suggested range.
It should be noted that AS5100.2-2004 [45] does not provide the length of the debris
mat in the flow direction (Ld ).
For the second approach [48], the debris modelling has been implemented by referring
to the regression curves obtained when experimenting on the jam formation of debris with
no uniform length on bridges with a single pier. The debris width, height and length are
calculated considering the following relationships:
 
Wd = 0.77 + 0.94 e−6.14 Fr L L L (8)
 
Hd = 0.39 − 0.46 e−5.77 Fr L L L (9)
 
Ld = 0.25 + 1.18 e−15.04 Fr L L L (10)

for 0.10 < Fr L < 0.40.


Wd , Hd , and Ld are the width,
p height, and length of the debris accumulations, respec-
tively (see Figure 6). Fr L = v/ gL L is the debris Froude number, g is the gravitational
acceleration, and L L is the so-called key log length. The key log is the key element that
triggers the accumulation of debris on the pier of the bridge and is represented by the
longest wood debris that can be encountered in the upstream reach of the river [48]. In
contrast to the recommendation of the Australian standard, the shape of the debris mat is
not rectangular but resembles a half-cone pointing downward (see Figure 6). In addition,
the size of the debris accumulation depends on the flow velocity.
spectively (see Figure 6). 𝐹𝑟𝐿 = 𝑣/√𝑔𝐿𝐿 is the debris Froude number, 𝑔 is the gravita-
tional acceleration, and 𝐿𝐿 is the so-called key log length. The key log is the key element
that triggers the accumulation of debris on the pier of the bridge and is represented by the
longest wood debris that can be encountered in the upstream reach of the river [48]. In
contrast to the recommendation of the Australian standard, the shape of the debris mat is
Water 2023, 15, 129 not rectangular but resembles a half-cone pointing downward (see Figure 6). In addition,11 of 29
the size of the debris accumulation depends on the flow velocity.

Figure
Figure 6. Idealized
6. Idealized sketch
sketch of a of a debris
debris accumulation
accumulation at theatbridge
the bridge pier according
pier according to Panici
to Panici and
and Al-
Almeida’s
meida’s modelmodel
[48]. [48].
FigureFigure adapted
adapted fromfrom the same
the same sourcesource
[48]. [48].

Equations (8)–(11) can only be solved after inputting the values of flow velocities
and assigning an appropriate key log length L L . The assessment of L L is not trivial, as
this would require persistent monitoring campaigns (i.e., aerial imagery) of the river and
riparian areas. As data on wood debris dynamics are missing, L L has to be selected from
one of the few datasets available in the literature. In particular, L L is selected as the 95%
percentile of the probability density function representing the datasets of debris lengths
observed in the South River (Virginia, United States) by [49]. The probability density
function that best fits the data is a log-normal distribution, which was also used by [48]
to reproduce the length distribution of the logs in the experiments. It is interesting to see
that the dataset published by [49] is somehow representative of the range of wood length
frequencies found in many rivers throughout the world (e.g., Chile, North Germany, Italy,
and New Zealand) [50]. Based on the above observation, the employed dataset is deemed to
be a good representative of the length of wood debris in general and can hence be used also
for the subject river. The value found for L L is 16 m, which complies well with the highest
range of heights of willow trees recorded in the basin to which the river belongs [51].
For the sake of clarity, Table 2 is introduced to outline the comparison between the
two debris modelling approaches. It is clear that, while for the first approach the debris
accumulation size is constant regardless of the hydraulic conditions, for the second ap-
proach, Wd , Ld , Hd vary with bulk flow characteristics and the length of the key log. Hence,
it is clear that the AS5100.2-2004 standard provides a more rigid and simplistic approach
compared to what is proposed by [48].

Table 2. Properties assigned to wood debris accumulation according to the two modelling approaches.

Wood Debris Accumulation Properties


shape Hd Wd Ld
Pier:
d =
W
AS5100.2-2004 rectangular 1.2–3 m 20 m not specified
min 1
2 ∑ SL
Deck : Wd = D L
Panici and Almeida’s
half conical f (v, g, H, L L ) f (v, g, H, L L ) f (v, g, H, L L )
model [48]

The magnitude of the drag force exerted by the debris jam is primarily a function of
the geometry of the debris accumulation and can be assessed through the empirical drag
equation (see Equation (5)). The drag force acting on the area of the pier underneath the
debris accumulation is also considered.
Water 2023, 15, 129 12 of 29

2.2.5. Local Scour Modelling


The depth of the local scour is computed according to the HEC-18 equation [52]:
"  0.35 #
H 0.43
ys = 2.0 K1 K2 K3 Fr a (11)
a

where K1 , K2 , and K3 are the correction factors for the pier nose shape, the angle of attack of
flow (θ), and riverbed conditions, respectively. a is the pier width, whereas Fr is the Froude
number and is defined as:
v
Fr = (12)
( g H )1/2
where v is the mean flow velocity upstream of the pier, g is the acceleration of gravity, and
H is the water height upstream of the pier. For round-nose piers with an angle of attack of
flow (θ ) equal to zero, the coefficients K1 and K2 both amount to 1.0. For clear-water scour,
K3 is equal to 1.1.
There is also value in considering the impact of debris on the local scour. The effect
of debris on the local scour of the pier is computed following the approach proposed by
Lagasse et al. [53]. The approach relies on the quantification of the equivalent pier width
(a∗d ), which is then used in Equation (11) instead of a to quantify the scour depth when
considering a debris jam. The equivalent pier width is defined as:

Kd1 ( Hd Wd ) ( Ld /H )Kd2 + ( H − Kd1 Hd ) a


a∗d = for Ld /H > 1.0 (13)
H
Kd1 ( Hd Wd ) + ( H − Kd1 Hd ) a
a∗d = for Ld /H ≤ 1.0 (14)
H
where Kd1 and Kd2 are factors depending on the shape of the debris raft (Kd1 = 0.79 and
Kd2 = −0.79 for rectangular debris raft; Kd1 = 0.21 and Kd2 = −0.17 for triangular debris
raft), and H is the depth of the approaching flow. Due to the lack of recommendations
in the Australian standard [45], Ld is set to be equal to H. This condition, according to
Equations (13) and (14) [53], allows the reproduction of the largest amplification of the
scour for each H − v considered.

Effect of Local Scour on SSI and Flood Loading


The action of the local scour on SSI and flood loading is modelled by reducing the
foundation embedded length (DE ) in Equations (1)–(4). Reducing DE leads to a lower
vertical, horizontal, rotational, and torsional stiffness of the foundation, which, in turn,
results in the higher flexibility of the pier of the bridge. As a result, a redistribution of the
flooding action between the central pier and the embankments and an additional settlement
of the central pier is ascertained.
In addition to the above-mentioned consequence, local scour may cause the exposure
of the foundation, thus leading to additional hydrodynamic loading. This effect is included
by assigning additional drag forces on the part of the foundation that is exposed due
to scour.

2.2.6. Parameters’ Uncertainty


The present study considers uncertainties in the bridge’s materials, geotechnical
parameters, hydrodynamic loads, local scour depths, and sizes of debris accumulation. The
variability of the model’s input parameters is introduced by multiplying the parameters
of the deterministic model by appropriately distributed random variables with a median
value of 1.0.
The statistical distributions of the random variables and their correlations are estab-
lished and computed based on the recommendations found in the literature (Table 3). Most
of the random variables are considered uncorrelated, except for the geotechnical character-
Water 2023, 15, 129 13 of 29

istics for which the correlation matrix shown in Table 4 is considered [54]. The concrete
elastic modulus is computed according to EN 1992-1-1 [40], and is therefore considered
to be fully correlated with the concrete strength. Similarly, the G of the soil is computed
considering a fixed Poisson’s ratio of 0.3. This results in a full correlation between G and E.

Table 3. Random variables used in the flood vulnerability analysis.

Parameters Probabilistic Distributions References


Concrete compressive strength ( f cm ) Normal (µ = 1, COV = 0.20) [55]
Steel yield strength ( f sy ) Log − normal (m
e = 1, COV = 0.07 ) [56]
Steel elastic modulus ( Es ) Log − normal (m
e = 1, COV = 0.03 ) [56]
Soil shear modulus ( G ) Log − normal (m
e = 1, COV = 0.30 ) [54,57]
Soil effective shear angle ( ϕ0 ) Normal (µ = 1, COV = 0.12) [54,57]
Soil unit weight (γ) Normal (µ = 1, COV = 0.10) [54]
Hydrodynamic drag forces ( Fd ) Normal (µ = 1, COV = 0.10) [24]
Local scour depth (ys ) Log − normal (m
e = 0.68, COV = 0.16 ) [58]
Debris width (Wd ) Normal (µ = 1, COV = 0.10) [48]
Debris height ( Hd ) Normal (µ = 1, COV = 0.25) [48]
Debris length ( Ld ) Normal (µ = 1, COV = 0.25) [48]

Table 4. Correlation matrix implemented for the soil characteristics taken from [54].

G ϕ0 γ
G 1.0 0.35 0.45
ϕ0 0.35 1.0 0.30
γ 0.45 0.30 1.0

2.2.7. Latin Hypercube Sampling


In this study, the Latin Hypercube Sampling (LHS) is operated following [59], whose
methodology is well known to be used for uncertainty analysis in seismic engineering
(e.g., [60–62]).
The output obtained from the LHS is a set of input model parameters, which defines a
set of deterministic numerical models. The set of models is assumed to fully describe the
statistical characteristics of the stochastic model. Thus, the propagation of the uncertainty
of the model’s inputs is achieved by analyzing the models within a set of deterministic
numerical analyses. The number of analyses equals the performed number of LHS sim-
ulations ( Nsim ). Within each analysis j (j = {1 . . . Nsim }), a flood analysis is performed to
assert whether a designated LS is exceeded or not. For this purpose, an index vector is
introduced for each LS and for each pair of H − v values:
 
0 i f LS is not exceed
ind LS j | H, v) = (15)
1 i f LS is exceed

The probability of exceeding a designated LS, P( LS| H, v), is computed as the ratio
between the number of simulations that leads to the exceedance of the designated LS
(NLS ( H, v)) and the total number of simulations Nsim :

NLS ( H, v)
P( LS| H, v) = (16)
Nsim
Water 2023, 15, 129 14 of 29

where NLS ( H, v) is calculated as the sum of the index vector as follows:


N
NLS ( H, v) = ∑ j=sim1 ind LS ( j | H, v) (17)

Selecting the optimal number of simulations (Nsim ) is a fundamental step. The required
number of simulations depends on the number of random input variables (Nvar ) and the
precision that is aimed to be achieved. Sensitivity analyses can help to obtain the optimal
number of simulations for the examined problem. An indication of the required Nsim can
be found based on the recommendation provided by [60], who suggests that, for practical
engineering applications, Nsim should be approximately 3timeslarger than the number of
input variables (Nvar ). In this study, Nsim is set to 40, which is 3.6-times larger than Nvar
(Nvar = 11). The selected Nsim produces a tolerance in the estimated LS’s probability of
1/40 = 0.025 = 2.5 %, which is considered to be sufficient for practical purposes.

3. Results
3.1. Model Validation
The performance of the proposed model was validated against measured modal peri-
ods from a modal survey. The modal survey involved an ambient vibration measurement
campaign with 12 tri-axial accelerometers placed on the bridge to measure its response to
vibrations from the environment. The results of the measurements were analyzed using
operational modal analysis, which allowed the identification of four translational modes of
the bridge’s deck. However, based on this campaign, it was not possible to obtain the global
translational modes of the bridge that would allow for more comprehensive validation of
the SSI modelling. An attempt was made to perform a lateral modal analysis with the use of
two long-stroke shakers exciting the bridge in the horizontal direction. Unfortunately, this
modal survey proved unsuccessful, as it was not possible to excite the bridge sufficiently to
obtain a reliable prediction of the bridge’s horizontal modes. The validation of the model
was thus based on best-available data, although further refinement/calibration could be
made if information on horizontal modes would be available.
The comparison of the calculated modal periods with the measured data is presented
in Table 5. Results indicated an excellent match between the calculated and measured
modal periods from the ambient vibration campaign.

Table 5. Comparison of the calculated and measured modal periods.

Period Period
No. Description
(s)—Calculated (s)—Measured
1 1st vertical mode of the deck 0.67 0.68
2 Global longitudinal mode 0.58 ND
3 2nd vertical mode of the deck 0.41 0.43
4 Global transverse mode 0.36 ND
5 Global vertical mode 0.21 ND
6 1st transverse mode of the deck 0.21 0.21
7 3rd vertical mode of the deck 0.18 0.19

Comparing mode shapes also indicates good agreement with on-site measurements,
though not presented for brevity.

3.2. Vulnerability Analysis


The vulnerability analysis is performed considering both Serviceability and Ultimate
limit states for three different scenarios. As an initial scenario, gravity, hydrodynamic,
buoyant loading, and local scour are taken into account. As part of the second and
third scenarios, debris loading is also included and modelled based on the AS5100.2-2004
standard and Panici and Almeida’s model [48], respectively.
Water 2023, 15, 129 15 of 29

3.2.1. Debris Accumulation Sizes


Figure 7 explores the information from Table 2 in more detail. The three panels in
Figure 7 display, respectively, the debris accumulation width, height and length as functions
of water height for the two different modelling approaches considered herein. With the
intention of representing the most significant outcomes, the results obtained with Panici and
Almeida’s approach [48] are plotted for values of water heights that represent the steady-
flow condition and the upper prediction bound. It can be seen that the two modelling
approaches diverge the most in the assessment of Wd . Specifically, Wd remains constant with
H in the Australian standard, whereas, for Panici and Almeida’s approach, Wd decreases
Water 2023, 15, x FOR PEER REVIEW 16 of 30
and it is also the parameter that is affected the most by the hydraulic conditions. A
similar outcome is found for Hd (Figure 7b). However, unlike Wd , Hd increases with
the water height when implementing Panici and Almeida’s approach. In Figure 7c it is
is interesting to notice
notice thatthat the
theimplementation
implementationofofthe thetwo
twodifferent
differentapproaches
approaches leads
leads to to
opposite
opposite trends
trends 𝐿𝑑L. dThis
in in . This
is is mainly
mainly due
due to to
thethefactfact that
that when
when representing
representing thethe AS5100.2-
AS5100.2-
2004
2004 guidelinesthe
guidelines valueofof𝐿L𝑑d is assumed. Overall,
thevalue Overall, ititcan
canbe
beobserved
observedthat,
that,inin
steady-flow
steady-
conditions,
flow Equations
conditions, Equations (8)–(10)
(8)–(10)lead to debris
lead to debris accumulations
accumulations more
moreextended
extended in in
width
width and
andlength compared
length compared to tothethe
ones
onesrecommended
recommended in in
thethe
Australian
Australianstandard.
standard. The
The opposite
opposite can
canbebesaid when
said when considering
considering thethe
case
case Hd𝐻
of of , where,
𝑑 , where, for higher
for values
higher of
values H
of and
𝐻 v
and(i.e.,
𝑣 upper
(i.e.,
prediction
upper bound)
prediction and and
bound) mostmostof the H examined,
of the 𝐻 examined, the the
height of the
height debris
of the accumulation
debris accumu-
envisioned in [48] overtakes the one proposed
lation envisioned in [48] overtakes the one proposed by [36]. by [36].

(a)

(b)
Figure 7. Cont.
Water 2023, 15, 129 16 of 29

(b)

(c)
Figure
Figure7. 7.
Relationship
Relationshipbetween
between(a)(a)the
thewidth,
width, (b) the height,
height, (c)
(c)the
thelength
lengthofofdebris
debris accumulation
accumulation and
and the water height (H) for the two different debris modelling approaches. The results are
the water height (H) for the two different debris modelling approaches. The results are represented repre-
sented forsteady
for the the steady flow condition
flow condition and forand
thefor the upper
upper prediction
prediction boundbound of the regression
of the regression curve
curve of of 3.
Figure
Figure 3.
3.2.2. Limit State Definition
The conditions for which serviceability and ultimate limit states apply are set based
on threshold conditions. The SLS is attained when the percentage of the foundation depth
uncovered by the local scour exceeds 50% (0.5L = 2.8 m), or when the water level reaches
the top part of the pier (H = 7.5 m). Due to the loading actions and potential accumulation
of debris, it is deemed to be too risky to define the SLS at a water height reaching the deck.
Hence, to ensure the operability of the bridge, routine inspections are expected to be carried
out to monitor the depth of the scour and the water heights. The ULS are met as soon as a
single component of the bridge fails. The following failure mechanisms are examined:
• Bending and shear failure at the base of the piers;
• Shear failure of the bearing above the pier and the abutments;
• Bending failure of the deck around the weak axis at the section above the piers and
the midspan of the bridge;
• Bearing failure of the foundation (vertical and bending capacity at the base);
• Failure due to local scour reaching the base of the foundation.
The calculation of the ultimate capacity of the structure (ULS) is defined based on the
unfactored capacity of the members. The bending capacity of the pier is conservatively
defined by the section modulus of the pier and the mean tensile strength of the concrete.
Note that the contribution of the reinforcement is neglected, as it is assumed that the amount
of reinforcement is below the minimum standard’s requirements [40]. The shear strength
of the pier is calculated according to equation 6.2a found in [40], which is applicable for
members without shear reinforcement. The shear capacities of the shear pin at the middle
bearing and the shear keys at the abutment bearing are computed considering the elastic
shear resistance indicated in EN-1993-1-1 [39]. The bending capacity of the deck is defined
based on the section modulus of the deck in the transverse direction (direction of the flow).
The failure of the foundation is assessed with the Meyerhof method [63]. The failure due to
local scour is conservatively assumed to be attained when the scour hole reaches the base
of the foundation, despite the fact that some capacity is likely to be maintained.

3.2.3. Results for Serviceability Limit State


Figures 8–10 illustrate the vulnerability surfaces obtained for the SLS and the three
scenarios (i.e., no debris load, debris load modelled based on AS5100.2-2004, and debris
load modelled according to Panici and Almeida’s model [48]). As can be seen from figures,
the vulnerability surfaces show two distinct regions: one below and one above the pier
height (H = 7.5 m). These regions arise as a result of the criteria that have been used for
the definition of the SLS. The region below H = 7.5 m represents the SLS of the bridge in
Water 2023, 15, x FOR PEER REVIEW 18 of 30
Water 2023, 15, x FOR PEER REVIEW 18 of 30

Water 2023, 15, 129 the 17 of 29


thevulnerability
vulnerabilityofofthe thebridge,
bridge,the theshape
shapeofofthe thewood
woodaccumulation
accumulationmay maybecome
becomemoremore
significant
significantthan thanitsitssize.
size.Despite
Despitethe thefact
factthat,
that,for
forsome
somehydraulic
hydraulicconditions,
conditions,wood
wooddebris
debris
sizes
sizesmaymaybebecomparably
comparablylarger largerthan
thanwhat
whatisisrecommended
recommendedininthe theAS5100.2-2004
AS5100.2-2004standard
standard
(see
(seeFigure
relation
Figureto 7),
7),Panici
the Paniciand
threshold andAlmeida’s
value method
set for
Almeida’s [48]
the scour
method [48]provides
depths less
(see
provides unfavorable
Section
less 3.2.2).results.
unfavorable On theThere-
results. other
There-
hand,
fore, the region above H = 7.5 m is the part of the surface for which the
fore, it is clear that, in certain circumstances, the bridge is more vulnerable based onthe
it is clear that, in certain circumstances, the bridge is more SLS
vulnerable of
basedthe bridge
on the
applies
shape in relation to the threshold valuethan
set for the water heights.
shapeofofthe thewood
wooddebris
debrisloading
loadingrather
rather thanitsitssize.
size.

(a)(a) (b)
(b)
Figure 8. (a) Contour and (b) 3-D surface plot of the flood vulnerability surface for the serviceability
Figure8.8.(a)
Figure (a)Contour
Contourand
and(b)(b)3-D
3-Dsurface
surfaceplot
plotofofthe
theflood
floodvulnerability
vulnerabilitysurface
surfacefor
forthe
theserviceability
serviceability
limit state (SLS) for the first scenario (no debris).
limitstate
limit state(SLS)
(SLS)for
forthe
thefirst
firstscenario
scenario(no
(nodebris).
debris).

(a)(a) (b)
(b)
Figure 9. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the serviceability limit
Figure 9. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the serviceability limit
Figure
state 9. (a)
(SLS) forContour andscenario
the second (b) 3-D surface plot of the
(AS5100.2-2004 flood fragility surface for the serviceability limit
standard).
state (SLS) for the second scenario (AS5100.2-2004 standard).
state (SLS) for the second scenario (AS5100.2-2004 standard).
Water 2023, 15, x FOR PEER REVIEW 19 of 30
Water 2023, 15, 129 18 of 29

(a) (b)
Figure 10. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the serviceability
Figure 10. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the serviceability
limit state (SLS) for the third scenario (Panici and Almeida’s model [48]).
limit state (SLS) for the third scenario (Panici and Almeida’s model [48]).

3.2.4.When
Results for Ultimate
comparing Limit
Figures Stateit is clear that, for high water levels (H > 7.5 m), the
8–10,
Figures
probability of11–13
exceedingrepresentthe SLSthe is
vulnerability
equal to 1.0surfaces
regardlessobtained for the ULS
of the scenario and the three
considered. On
scenarios.
the When
other hand, forcomparing
values of HFigure < 7.5 m, 11 itwith Figurethat
is evident 12, the
andcontributions
Figure 11 with Figure
of the local13, it is
scour
and
clearthe mutual
that interaction
the debris betweenincreases
accumulation the localthescour
ULSand debris increase
exceedance the likelihood
probability significantly of
reaching
compared thetoSLS. In contrast
the scenario to the first
without scenario,
debris. In this where local ascour
latter case, has minimal
moderate percentage impact on
of fail-
the
ureresults, the second
is expected to occur scenario
within hasthetheupper
highest probability
bound of the of𝐻 exceeding the SLS.and
− 𝑣 relationship In this
only case,
for
the SLS is very likely to be exceeded for water heights as low as
extreme floods (i.e., when the deck is fully submerged) (Figure 11). The use of the stand-3 m (not nearly half of the
pier’s
ard’s height)
modelling andapproach
flow velocitiesresults asinlowtheaslargest
0.50 m/s.
ULSInexceedance
the third scenario, the vulnerability
probabilities (Figure 12),
surface
while the indicates
results aobtained
lower probability of exceeding
when implementing theand
Panici SLSAlmeida’s
comparedapproach
to the second[48] showone
and
thatitthe
is mainly
bridge occurring
failure is only in the area located
expected in the
to occur upper
within bound
the the H − v bound
upperofprediction relationship.
of the
This
𝐻 − discrepancy
𝑣 relationship between
(Figurethe 13).results of the second and third scenarios results from the
different extent of local
A detailed examination of scour and, additionally,
Figures 12 and 13 from adopting
suggests that different
both debris approaches
models yield for
modelling the shape of the wood accumulation. Indeed, when assessing
the largest ULS exceedance probability when the water heights reach the deck (i.e., prob- the vulnerability of
the bridge, the shape of the wood accumulation may become more
ability values of approximately 80% for 𝐻 > 8.15). In such a condition, the debris loading significant than its size.
Despite
is assumedthe fact
to actthat,
over forthesome hydraulic
entire length of conditions,
the deck, wood debris sizes
and therefore may be comparably
the hydrodynamic load-
larger than what is recommended in the AS5100.2-2004 standard
ing on the bridge reaches its maximum value. For water heights lower than the bridge (see Figure 7), Panici
and
deck,Almeida’s
the failure method
of the[48] provides
bridge appears lessasunfavorable
an unlikelyresults.
event Therefore, it is clearthe
when considering that, in
third
certain circumstances, the bridge is more vulnerable based on the
scenario. On the other hand, for the second scenario, a non-negligible failure probability shape of the wood debris
loading
occurs forrather
water than its size.
heights approximately equal to 4 m. This is owing to the scour reaching
the bottom of the foundation by triggering the failure of the bridge. Such a failure mech-
3.2.4. Results for Ultimate Limit State
anism is not observed when considering the results obtained with Panici and Almeida’s
modelFigures
[48], 11–13
whichrepresent
is found the vulnerability
to lead to a less surfaces obtainedcompared
severe scouring for the ULS to and
the the
other three
ap-
scenarios.
proach. Hence, even though the area of the debris accumulation in [48] is often bigger,itthe
When comparing Figure 11 with Figure 12, and Figure 11 with Figure 13, is
clear that the debris accumulation increases the ULS exceedance probability significantly
shape factors from Equations (13) and (14) show that the shape of debris is the leading
compared to the scenario without debris. In this latter case, a moderate percentage of failure
variable when computing the amplification of local scour. Furthermore, the freedom al-
is expected to occur within the upper bound of the H − v relationship and only for extreme
lowed in the AS5100.2-2004 standard of assigning suggested and custom values to the size
floods (i.e., when the deck is fully submerged) (Figure 11). The use of the standard’s
and shape of debris accumulations has a relevant role in the overall SLS and ULS exceed-
modelling approach results in the largest ULS exceedance probabilities (Figure 12), while
ance probabilities.
the results obtained when implementing Panici and Almeida’s approach [48] show that the
bridge failure is only expected to occur within the upper prediction bound of the H − v
relationship (Figure 13).
Water 2023, 15, x FOR PEER REVIEW 20 of 30
Water2023,
Water 15,x129
2023,15, FOR PEER REVIEW 2019ofof30
29

(a)
(a) (b)
(b)
Figure 11. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the ultimate limit
Figure
Figure11. (a)Contour
11.(a) Contourand and(b)
(b)3-D
3-Dsurface
surfaceplot
plotofofthe
theflood
floodfragility
fragilitysurface
surfacefor
forthe
theultimate
ultimatelimit
limit
state (ULS) for the first scenario (no debris).
state
state(ULS)
(ULS)for
forthe
thefirst
firstscenario
scenario(no
(nodebris).
debris).

(a)
(a) (b)
(b)
Figure 12. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the ultimate limit
Figure 12. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the ultimate limit
Figure
state (a) Contour
12. for
(ULS) and
the second (b) 3-D(AS5100.2-2004
scenario surface plot ofstandard).
the flood fragility surface for the ultimate limit
state (ULS) for the second scenario (AS5100.2-2004 standard).
state (ULS) for the second scenario (AS5100.2-2004 standard).
Water
Water 2023,
2023, 15,
15, x129
FOR PEER REVIEW 2120ofof30
29

(a) (b)
Figure 13. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the ultimate limit
Figure 13. (a) Contour and (b) 3-D surface plot of the flood fragility surface for the ultimate limit
state (ULS) for the third scenario (Panici and Almeida’s model [48]).
state (ULS) for the third scenario (Panici and Almeida’s model [48]).

3.3. Sensitivity
A detailedAnalysis
examination of Figures 12 and 13 suggests that both debris models yield the
This
largest ULSsection examines
exceedance the extent
probability to which
when the scour
the water heightsdepth
reachis the
affected by implement-
deck (i.e., probability
ing the of
values three scenarios (Section
approximately 80% for 3.3.1).
H >The bridge’s
8.15). In such failure utilization
a condition, theratios
debris areloading
also ex-is
plored
assumed (Section 3.3.2).the
to act over Note thatlength
entire all theofanalyses
the deck,areand only implemented
therefore for the upper
the hydrodynamic per-
loading
on the bridge
centile bound reaches its maximum
(95th percentile) of the 𝐻 −For
value. 𝑣 relationship,
water heightsfor lower than
which thethechance
bridgeofdeck, the
bridge
failure is
failure ofthe
themost
bridge appears as an unlikely event when considering the third scenario. On
pronounced.
the other hand, for the second scenario, a non-negligible failure probability occurs for water
heights
3.3.1. approximately
Scour Depths equal to 4 m. This is owing to the scour reaching the bottom of the
foundation
In Figure 14, the 5th, the
by triggering 50th failure of the
and 95th bridge. Such
percentiles a failure
of the scour mechanism is not observed
depths are plotted against
when considering the results obtained with Panici and Almeida’s
the water heights for the three scenarios. The values of the scour depths are computed model [48], whichbyis
found to lead to
implementing 40aLHS
less severe scouring
simulations compared
at each to the other
water height, approach.
considering theHence,
upper even
bound though
𝐻−
the area of the debris accumulation in [48] is often bigger, the
𝑣 relationship. The scour depths at which the SLS and ULS are reached are also repre-shape factors from Equations
(13) and
sented (14)horizontal
with show that dotted
the shape of debris
lines. As canisbethe leadingregardless
noticed, variable whenof thecomputing
percentile the am-
exam-
plification of local scour. Furthermore, the freedom allowed in the AS5100.2-2004 standard
ined, the largest scour depths are obtained for the second scenario, whereas the lowest
of assigning suggested and custom values to the size and shape of debris accumulations
values are attributed to the first scenario. The results obtained with Panici and Almeida’s
has a relevant role in the overall SLS and ULS exceedance probabilities.
model [48] mediate the extreme values of scours obtained for the aforementioned cases.
When debris accumulation
3.3. Sensitivity Analysis is modelled according to the AS5100.2-2004 standard, the
higher amplification of the scour occurs at lower water heights. This phenomenon is less
This section examines the extent to which the scour depth is affected by implementing
pronounced when implementing the other debris model, for which the amplification of
the three scenarios (Section 3.3.1). The bridge’s failure utilization ratios are also explored
the scour is more uniform over different water heights. As mentioned, this is likely a con-
(Section 3.3.2). Note that all the analyses are only implemented for the upper percentile
sequence of a more favorable shape of the debris accumulation (i.e., triangular) that also
bound (95th percentile) of the H − v relationship, for which the chance of bridge failure is
results in smaller scour depths [56].
the most pronounced.
For the first scenario, the SLS limit state is expected to be exceeded only for a few of
the most unfavorable
3.3.1. Scour Depths LHS simulations with the largest scour depths (95th percentile). This
offers an explanation of the low SLS exceedance probabilities shown in Figure 8 for water
In Figure 14, the 5th, 50th and 95th percentiles of the scour depths are plotted against
height
the waterbelow 𝐻 = 7.5
heights form.theOn the other
three hand,The
scenarios. the high
valuesSLS ofexceedance probability
the scour depths obtained
are computed
for the second scenario (Figure 9) can be explained by SLS
by implementing 40 LHS simulations at each water height, considering the upper boundscour depth limit being ex-
ceeded even for LHS simulations with the lowest scour predictions
H − v relationship. The scour depths at which the SLS and ULS are reached are also (5th percentile). For
the third scenario,
represented the high dotted
with horizontal SLS exceedance probabilities
lines. As can be noticed, occurring
regardless forofwater heights
the percentile
above
examined,3 m (Figure
the largest10) are consistent
scour depths are withobtained
the results
for illustrated
the secondinscenario,
Figure 14b,c,
whereaswhich
the
lowest values are attributed to the first scenario. The results obtained with Panici and
Water 2023, 15, 129 21 of 29

Almeida’s model [48] mediate the extreme values of scours obtained for the aforementioned
cases. When debris accumulation is modelled according to the AS5100.2-2004 standard,
Water 2023, 15, x FOR PEER REVIEW the higher amplification of the scour occurs at lower water heights. This phenomenon 22 of 30 is
less pronounced when implementing the other debris model, for which the amplification
of the scour is more uniform over different water heights. As mentioned, this is likely a
indicate that theofSLS
consequence limit favorable
a more is exceededshape
only for
of values of scour
the debris depths above
accumulation the triangular)
(i.e., 50th per- that
centile.
also results in smaller scour depths [56].

(a)

(b)

(c)
Figure
Figure14.
14.Representation of the
Representation (a) 5th
of the (a) percentile, (b) 50th
5th percentile, (b)percentile and (c)and
50th percentile 95th(c)
percentile of scour of scour
95th percentile
depths (𝑦𝑠 ) against the water heights (𝐻) for the three different scenarios and the upper prediction
depths (ys ) against the water heights (H) for the three different scenarios and the upper prediction
bound of the 𝐻 − 𝑣 relationship.
bound of the H − v relationship.
Water 2023, 15, 129 22 of 29

For the first scenario, the SLS limit state is expected to be exceeded only for a few
of the most unfavorable LHS simulations with the largest scour depths (95th percentile).
This offers an explanation of the low SLS exceedance probabilities shown in Figure 8 for
water height below H = 7.5 m. On the other hand, the high SLS exceedance probability
Water 2023, 15, x FOR PEER REVIEW obtained for the second scenario (Figure 9) can be explained by SLS scour depth limit 23 being
of 30
exceeded even for LHS simulations with the lowest scour predictions (5th percentile). For
the third scenario, the high SLS exceedance probabilities occurring for water heights above
3mBy(Figure 10) are
comparing theconsistent
results inwith the results
Figure illustrated
14, it can in Figure
be observed that 14b,c, which
the ULS indicate
is only ex-
that the SLS limit is exceeded only for scour depths above the 50th percentile.
ceeded in the second scenario and the 95th percentile scour-depth prediction (Figure 14c).
By comparing the results in Figure 14, it can be observed that the ULS is only exceeded
This tendency supports the non-zero ULS exceedance probabilities shown in Figure 12 for
in the second scenario and the 95th percentile scour-depth prediction (Figure 14c). This
water heights below 8.15 m.
tendency supports the non-zero ULS exceedance probabilities shown in Figure 12 for water
heights below 8.15 m.
3.3.2. Utilization Ratios for ULS
Figure
3.3.2. 15 illustrates
Utilization Ratios the variation of the failure utilization ratio (𝑈𝑅) of the bridge
for ULS
(5th, 50th, and 95th percentiles) against the
Figure 15 illustrates the variation water
of the heights
failure for the three
utilization scenarios
ratio (UR) of theand the
bridge
upper
(5th,prediction bound
50th, and 95th of the 𝐻 against
percentiles) − 𝑣 relationship. The utilization
the water heights ratioscenarios
for the three of the bridge is
and the
defined
upperasprediction
the maximum bound ratio between
of the H − vthe demand andThe
relationship. capacity of theratio
utilization individual mem-
of the bridge
bersis of the bridge.
defined as the Amaximum
value of ratio
utilization
between ratio
thelarger
demand thanand 1 indicates
capacity the failure
of the of the
individual
bridge (exceedance
members of the ULS
of the bridge. as defined
A value in Section
of utilization ratio3.2.2).
largerThethan5th, 50th and the
1 indicates 95thfailure
percen-
of
the
tiles bridge (exceedance
utilization ratios of the of bridge
the ULSare as obtained
defined in bySection 3.2.2).40The
performing LHS5th, 50th and 95th
simulations for
eachpercentiles utilization
water height ratios of
and related the bridge are
upper-bound obtainedInby
velocity. the performing 40 LHS
scenario with no simulations
debris, the
for each water
exceedance of theheight
ULS ofandtherelated
bridgeupper-bound
is only expectedvelocity. In thefor
to occur scenario
waterwith no debris,
heights above the
10
m and for a value of utilization ratio above the 50th percentile. This condition results in m
exceedance of the ULS of the bridge is only expected to occur for water heights above 10 a
and probability
failure for a value of utilization
equal ratio50%,
to roughly above the 50th
which percentile.
is consistent withThis
thecondition
results inresults
Figurein11.a
failure
In the caseprobability
when debrisequal to roughly 50%,
accumulation which is according
is modelled consistent to with the results instandard,
AS5100.2-2004 Figure 11.
the exceedance of the ULS is expected over the entire range of water heights forstandard,
In the case when debris accumulation is modelled according to AS5100.2-2004 the 95%
the exceedance of the ULS is expected over the entire range of water heights for the 95%
percentile of utilization ratio, reaching particularly high values when the water height
percentile of utilization ratio, reaching particularly high values when the water height
reaches the deck (𝐻 > 8.15 m). This outcome is in line with what is shown in Figure 12, in
reaches the deck (H > 8.15 m). This outcome is in line with what is shown in Figure 12, in
which the non-zero ULS exceedance probabilities are reached over the entire upper part
which the non-zero ULS exceedance probabilities are reached over the entire upper part of
of the vulnerability surface.
the vulnerability surface.

(a)

Figure 15. Cont.


Water
Water 2023,
2023, 15,
15, x129
FOR PEER REVIEW 2423ofof30
29

(b)

(c)
Figure
Figure 15.
15. Representation
Representationofofthe
the(a)
(a)5th,
5th,(b)
(b) 50th,
50th, and
and (c)
(c) 95th
95th percentiles
percentiles of
of the
the failure
failure utilization
utilization
ratios (𝑈𝑅) against the water heights (𝐻) for the three different scenarios and for the
ratios (UR) against the water heights (H) for the three different scenarios and for the upperupper predic-
prediction
tion bound of the 𝐻 − 𝑣 relationship.
bound of the H − v relationship.

On
On the
the other
other hand,
hand, when
when modelling
modelling the the debris
debris loading
loading with
with Panici
Panici and
and Almeida’s
Almeida’s
model
model[48],
[48],the
the50th
50thand
and95th
95thpercentiles
percentilesofofutilization
utilizationratios
ratiosexceed
exceedthetheULS
ULSlimit
limitonly
onlyforfor
water
water heights
heights above
above 8.15
8.15 m,
m, thus
thus explaining
explaining the the results
results from
from Figure
Figure 13.
13.
Assessing
Assessing thethe maximum
maximum values values of 𝑈𝑅 across
of UR acrossthe
theULS
ULSsimulations
simulationsenables
enablesthe theidentifi-
iden-
tification of the leading mechanism that threatens the bridge’s performance
cation of the leading mechanism that threatens the bridge’s performance (i.e., the mecha- (i.e., the mech-
anism with the
nism with the largest 𝑈𝑅). In
largest UR). In this
this context,
context, the
the most
most critical
critical mechanisms
mechanisms are are identified
identifiedas as
the
thelocal
localscour
scourandandthethefailure
failureofofthe
thepier-bearing
pier-bearingsystem.
system.Figure
Figure1616displays
displaysthethefrequency
frequency
of
of occurrence
occurrence of of such
such mechanisms
mechanisms withinwithin40 40 LHS
LHS simulations
simulationsfor for the
the different
different scenarios
scenarios
and
and within
within the upper
upper prediction
predictionbound
boundofofthetheH𝐻−−v 𝑣relationship.
relationship. Overall,
Overall, regardless
regardless of
of the
the scenario
scenario considered,
considered, the the
locallocal
scourscour proves
proves to betothe
bepredominant
the predominant mechanism
mechanism (i.e., (i.e.,
with
with maximum
maximum utilization
utilization ratio) ratio)
withinwithin
a rangeaof range
waterofheight
waterbetween
height between
2 and 8.15 2 m.
andWhen8.15 m.
the
When
water the water
reaches thereaches
bridgethe bridge
deck, deck,
the loss of the
the loss of the
bridge’s bridge’s performance
performance is predom-
is predominantly asso-
inantly associated
ciated with with a compromised
a compromised functionalityfunctionality of the pier-bearing
of the pier-bearing system.however,
system. The results, The re-
sults,
differhowever,
depending differ depending
on the scenarioon the scenario
considered. considered.
Data reveal thatData reveal
when thatdebris
wood when woodforces
are neglected,
debris forces are theneglected,
potentialthe of the deck unseating
potential of the deck due to the failure
unseating due toofthethefailure
pier bearing
of the
Water 2023,
Water 2023, 15,
15, 129
x FOR PEER REVIEW 25
24 of 30
of 29

(Figure 16a) is
pier bearing less pronounced
(Figure compared tocompared
16a) is less pronounced Figure 16b,c, in which
to Figure a in
16b,c, higher
which frequency
a higher
of bearing of
frequency failure is obtained
bearing failure isat lower water
obtained height.
at lower waterAtheight.
the same time,
At the it istime,
same clearitthat the
is clear
implementation of two different
that the implementation wood debris
of two different woodmodels
debrishas a minor
models hasimpact on impact
a minor the frequency
on the
of occurrence
frequency of both mechanisms.
of occurrence of both mechanisms.

No debris AS5100.2-2004 standard Panici and Almeida’s model

(a) (b) (c)


Figure 16.
Figure 16. Frequency
Frequency of
of occurrence
occurrence ofof two
two leading
leading mechanisms
mechanisms (i.e.,
(i.e., local
local scour
scour and
and pier
pier bearing
bearing
failure) within LHS simulations for the (a) first, (b) second, and (c) third debris scenario and the
failure) within LHS simulations for the (a) first, (b) second, and (c) third debris scenario and the
upper prediction bound of the 𝐻 − 𝑣 relationship. Note that the frequencies between 𝐻 = 2 m and
upper prediction bound of the H − v relationship. Note that the frequencies between H = 2 m and
𝐻 = 8 m are omitted, since, in this range, local scour is the leading mechanism.
H = 8 m are omitted, since, in this range, local scour is the leading mechanism.
4. Discussion
4. Discussion
In this
In thispaper,
paper,a afloodflood vulnerability
vulnerability analysis
analysis of a of a real-case
real-case roadway roadway
bridgebridge is per-
is performed.
formed. The vulnerability of the bridge is examined by considering
The vulnerability of the bridge is examined by considering the concurrent action of gravity the concurrent action
of gravity loads, hydrodynamic loads, wood debris accumulation
loads, hydrodynamic loads, wood debris accumulation and local scour. The interaction and local scour. The
interaction
between between
local scour and localthescour and the accumulation
accumulation of debris is also of debris is alsoand
considered, considered, and its
its implementa-
implementation
tion is based on an is based on an equivalent-pier-width
equivalent-pier-width approach. Due approach.
to the lack Dueofto the lack of
stationary station-
hydraulic
ary hydraulic
conditions conditions
at the site, the at the site, theanalysis
vulnerability vulnerability analysiswith
is performed is performed
consideration with ofconsid-
both
flow velocity and water height as input parameters. This leads to the representation ofrep-
eration of both flow velocity and water height as input parameters. This leads to the the
resentation ofanalysis
vulnerability the vulnerability
results as aanalysis
surface in results as a surface in domain.
a three-dimensional a three-dimensional do-
main.The main purpose of this paper is to examine the results of three different scenarios.
One The main purpose
examines of thiswith
the condition papernois debris,
to examine while thetheresults
second of three different
and third scenarios.
evaluate two
One examines
different debristhe condition
modelling with no debris,
approaches. The firstwhile the second
approach and
relies onthird evaluate two dif-
the recommendations
ferent debris
drawn from the modelling approaches.
AS5100.2-2004 standard, Thewhereas
first approach
the second relies on the refers
approach recommendations
to the model
drawn from the AS5100.2-2004 standard, whereas the
proposed by Panici and Almeida [48], which, compared to the standards-based second approach refers to methodol-
the model
proposed
ogy, involvesby aPanici
more and Almeida
elaborate and [48],
refinedwhich, compared
analytical to the The
procedure. standards-based
research of themethod-second
ology, involves
approach a moreby
is motivated elaborate
the absence and of refined
accurateanalytical
guidelines procedure. The research
for modelling processesof the
of
wood
seconddebris accumulation
approach is motivated at bridge
by the piers.
absence of accurate guidelines for modelling pro-
cessesBased on the
of wood results,
debris it is clear that
accumulation the bridge
at bridge is more likely to exceed its serviceability
piers.
and ultimate
Based onlimit states in
the results, it ispresence
clear thatofthe wood debris
bridge compared
is more likely totoexceed
the condition with no
its serviceability
debris. The accumulation of debris at the bridge pier
and ultimate limit states in presence of wood debris compared the condition with is shown to be responsible for the
no
increase
debris. Thein local scour depths.
accumulation of debrisAs the at debris
the bridge and pier
scourisaction
showninteract, the SLS and
to be responsible forULS
the
of the bridge
increase occur
in local scourfordepths.
smallerAs flooding
the debriseventsandcompared
scour action to the scenario
interact, withand
the SLS no ULS
debris.
of
Nevertheless,
the bridge occur the for
bridge is more
smaller likelyevents
flooding to fail for values of
compared to H
theand v located
scenario with in no
thedebris.
upper
prediction
Nevertheless,interval of the H
the bridge is− v relationship.
more likely to fail This
forfinding
values is ofnot
𝐻 dependent
and 𝑣 located on the in modelling
the upper
approach considered. Overall, the results of this study indicate
prediction interval of the 𝐻 − 𝑣 relationship. This finding is not dependent on the mod- that the use of Panici and
Almeida’s model [48] offers less conservative estimates of the bridge
elling approach considered. Overall, the results of this study indicate that the use of Panici response to severe
floodings.
and Almeida’sThis may model be due
[48] to the fact
offers lessthat the novel modelling
conservative estimates approach
of the bridge is ableresponse
to provideto
Water 2023, 15, 129 25 of 29

a more analytical definition of the size and shape of the debris accumulation. The triangular
shape, in fact, yields less severe scouring depths compared to the use of the rectangular
shape suggested in the AS5100.2-2004 standard [53].
A sensitivity analysis is also performed and is used to examine the variation of the
scour depths, bridge’s utilization ratios and governing failure mechanisms for the three
scenarios and different water heights. From this analysis, it is shown that the AS5100.2-
2004 standard produces the most unfavorable results. For instance, the exceedance of the
ultimate limit states due to scour only occurs when implementing the recommendations
provided in the Australian Standard. A similar outcome occurs when examining the
values of utilization ratios of the structure. It is clear that for most of the values of H, the
AS5100.2-2004 Standard is the only scenario that leads to the exceedance of the ULS for
the highest values of the utilization ratio (95% percentile). It also emerges that local scour
plays a crucial role in most of the hydraulic conditions considered, and its influence is
more significant for the second scenario than the third. This confirms what was found
in the literature [64] and makes it clear that computing the magnitude of the local scour
is essential when examining bridge vulnerability to flood and bridge failure modes. For
water height reaching the deck, the performance of the bridge is mainly threatened by the
failure of the pier bearing instead. However, in this case, the difference between the two
debris models becomes less evident.
In spite of the effectiveness of the presented vulnerability approach, there are limi-
tations that need to be taken into account. For instance, the poor quality and amount of
hydraulic data and a lack of an appropriate numerical modelling approach do not allow a
reasonable estimation of the hydrodynamic-related loads. Hydraulic data input are essen-
tial in flood vulnerability analysis. Hence, more appropriate and consistent monitoring of
the hydraulic variables is encouraged as it can be extremely beneficial in the vulnerability
assessment of riverine bridges. Similarly, the knowledge gap related to the mechanisms
governing the accumulation of debris jams at bridge piers, as well as the lack of monitoring
data on wood debris recruitment and transportation represents a limitation for the reliabil-
ity of the results of the analysis. Another aspect to consider is the simplified finite-element
modelling employed in this study, which, due to a complicated interaction between the
structure, soil, hydrodynamic loads and local scour, owns further examinations. A similar
evaluation applies to the coefficient of variations of the materials, geotechnical characteris-
tics, hydrodynamic forces, local scour depth and size of wood debris accumulation. The
values assigned to the coefficient of variations are, in fact, based on isolated recommenda-
tions found in the literature. Moreover, important processes such as afflux, river bed-load
transport and the degradation of the bridge’s materials are not included in the analysis
as the present study does not account for the time-dependent assessment of the bridge’s
vulnerability. Due to the complex interplay between debris loads, scour, and hydraulic
forces, the research examination assessing bridge flood vulnerability should be carefully
addressed. Despite the acknowledged limitations, the presented framework bridges some
of the gaps existing in the standards-based approach providing a more up-to-date and
faithful methodology for bridge vulnerability analysis. Therefore, more effort should be
devoted to improving and updating the existing standards and guidelines concerning
the modelling of hydrodynamic-related processes. Particularly, the interplay between
debris accumulation and local scour would need further examination as it is shown to
substantially affect the bridge’s vulnerability.
It is hoped that this study will inspire future research into improving and upgrad-
ing the vulnerability analysis framework. For instance, the implementation of faithful
hydraulic numerical modelling techniques, along with local scour monitoring solutions,
can help to validate the proposed modelling approach and assess the performance of the
asset throughout its lifespan. At the same time, the developed approach represents a
valuable tool for supporting targeted visual inspection and Structural Health Monitoring
campaigns. Vulnerability surfaces can provide useful insight into the selection of the key
parameters that should be verified and monitored at different stages of a bridge’s life-cycle
Water 2023, 15, 129 26 of 29

to determine its level of performance [65]. Furthermore, considering a life-cycle perspec-


tive, bridge vulnerability assessment is also beneficial for quantitatively determining the
risk severity of hazards on bridges and helping to evaluate the effectiveness of potential
resilient countermeasures for risk mitigation. This study can provide policymakers, engi-
neering consultants, bridge engineers and stakeholders with a valid and reliable method
of estimating the anticipated economic and functional losses associated with flooding
hazards, implementing effective decision-making processes, and selecting sustainable and
resilient strategies.

Author Contributions: Conceptualization, methodology M.K., A.A. and V.B.; software, validation,
formal analysis, M.K.; investigation, resources, data curation M.K., A.A. and V.B.; writing—original
draft preparation, M.K.; writing—review and editing, A.A. and V.B.; visualization, M.K. and A.A.
and V.B.; project administration, A.A.; funding acquisition, A.A. All authors have read and agreed to
the published version of the manuscript.
Funding: This research was funded by the European Union Civil Protection Mechanism, under
the UCPM-2019-PP-AG call, Grant Agreement Number 874421, oVERFLOw project (Vulnerability
assessment of embankments and bridges exposed to flooding hazard). The authors would also like to
express their gratitude for the support received from the Slovenian Research Agency: research core
funding No. P2-0273 and from the infrastructure program: No. I0-032.
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.

List of Symbols and Notations

a Pier width
a∗d Equivalent pier width
Ad Projected wetted area in the direction of the flow
Al Wetted area perpendicular to the flow
c0 Effective cohesion of the soil
Cd Drag coefficient
Cl Lift coefficient
CPT Cone Penetration Test
D Diameter of a well foundation
DE Embedment depth of the foundation
DL Length of the deck
dsp Wetted depth of the deck
dss Wetted depth of the solid superstructure
dwgs Distance from the girder soffit to the flood water surface
E Soil elastic modulus
Es Steel elastic modulus
f cm Concrete compressive strength
f sy Steel yield strength
Fd Hydrodynamic drag force
Fl Hydrodynamic lift force
FrL Debris Froude number
Fr Froude number
g Gravitational acceleration
G Shear modulus of the soil
H Water height
Hd Height of wood debris accumulation
IM Intensity measure
Water 2023, 15, 129 27 of 29

ind LS ( j| H, v) Index vector for each LS and each H-v


j Index of j-th LHS simulation
K1 Correction factor for the pier nose shape
K2 Correction factor for the angle of attack flow
K3 Correction factor for the riverbed condition
Correction factors used in the calculation of the equivalent pier width depending
Kd,1 Kd,2
on the shape of the debris raft
K eH Horizontal embedded stiffness of the well foundation
KRe Rocking embedded stiffness of the well foundation
KVe Vertical embedded stiffness of the well foundation
KTe Torsional embedded stiffness of the well foundation
L Depth of the foundation
Ld Length of wood debris accumulation in the flow direction
LL Key log length
LHS Latin Hypercube Sampling
LS Limit state
N60 Equivalent number of Standard Penetration Test (SPT) blows of the soil
NLS ( H, v) Number of simulations that leads to the exceedance of LS for a given H-v
Nsim Number of LHS simulations
Nvar Number of random input variables
P( LS| H, v) Probability of exceeding the designated LS for a given H-v
PI Prediction Intervals
Pr Proximity ratio
Q Water discharge
RC Reinforced Concrete
Rf Foundation radius
SL Span length of the bridge
Sr Relative submergence of the deck
SLS Serviceability Limit State
SSI Soil–Structure Interaction
ULS Ultimate Limit State
UR Utilization ratio
v Mean flow velocity
Wd Width of wood debris accumulation
y gs Average vertical distance from the girder soffit to the riverbed
ys Local scour depth
ϕ0 Effective shear angle of the soil
γ Soil unit weight
ν Poisson’s ratio
θ Angle of flow attack

References
1. Jean-Louis, J.B.; Paolo, G.; Congpu, Y. Statistical, Risk, and Reliability Analyses of Bridge Scour. J. Geotech. Geoenviron. Eng. 2014,
140, 4013011. [CrossRef]
2. Benn, J. Railway bridge failure during flooding in the UK and Ireland. Proc. Inst. Civ. Eng. Forensic Eng. 2013, 166, 163–170.
[CrossRef]
3. Kumalasari, W.; Hadipriono, F.C. Analysis of Recent Bridge Failures in the United States. J. Perform. Constr. Facil. 2003, 17,
144–150. [CrossRef]
4. Zampieri, P.; Zanini, M.A.; Faleschini, F.; Hofer, L.; Pellegrino, C. Failure analysis of masonry arch bridges subject to local pier
scour. Eng. Fail. Anal. 2017, 79, 371–384. [CrossRef]
5. Stefanidis, S.; Alexandridis, V.; Theodoridou, T. Flood exposure of residential areas and infrastructure in Greece. Hydrology 2022,
9, 145. [CrossRef]
6. Karatzetzou, A.; Stefanidis, S.; Stefanidou, S.; Tsinidis, G.; Pitilakis, D. Unified hazard models for risk assessment of transportation
networks in a multi-hazard environment. Int. J. Disaster Risk Reduct. 2022, 75, 102960. [CrossRef]
7. Banerjee, S.; Shinozuka, M. Nonlinear Static Procedure for Seismic Vulnerability Assessment of Bridges. Comput. Civ. Infrastruct.
Eng. 2007, 22, 293–305. [CrossRef]
8. Ma, X.; Zhang, W. Dynamic amplification responses of short span bridges considering scour and debris impacts. Eng. Struct.
2022, 252, 113644. [CrossRef]
Water 2023, 15, 129 28 of 29

9. Pagliara, S.; Carnacina, I. Influence of wood debris accumulation on bridge pier scour. J. Hydraul. Eng. 2011, 137, 254–261.
[CrossRef]
10. Prendergast, L.J.; Gavin, K. A review of bridge scour monitoring techniques. J. Rock Mech. Geotech. Eng. 2014, 6, 138–149.
[CrossRef]
11. Ebrahimi, M.; Djordjević, S.; Panici, D.; Tabor, G.; Kripakaran, P. A method for evaluating local scour depth at bridge piers due to
debris accumulation. In Proceedings of the Institution of Civil Engineers—Bridge Engineering; Thomas Telford Ltd.: London,
UK, 2020; Volume 173, pp. 86–99.
12. Shen, H.W.; Schneider, V.R.; Karaki, S. Local scour around bridge piers. J. Hydraul. Div. 1969, 95, 1919–1940. [CrossRef]
13. Lagasse, P.F. Countermeasures to Protect Bridge Piers from Scour; Transportation Research Board: Washington, DC, USA, 2007;
Volume 593.
14. Lin, C.; Han, J.; Bennett, C.; Parsons, R.L. Case history analysis of bridge failures due to scour. In Climatic Effects on Pavement and
Geotechnical Infrastructure; American Society of Civil Engineers: Reston, VA, USA, 2014; pp. 204–216.
15. Zhang, Y. Economic Impact of Bridge Damage in a Flood Event. Master‘s Thesis, RMIT University, Melbourne, Australia, 2016.
16. Stewart, M.G. Reliability-based assessment of ageing bridges using risk ranking and life cycle cost decision analyses. Reliab. Eng.
Syst. Saf. 2001, 74, 263–273. [CrossRef]
17. Billah, A.; Alam, M.S. Seismic fragility assessment of highway bridges: A state-ofthe-art review. Struct. Infrastruct. Eng. 2015, 11,
804–832. [CrossRef]
18. Karamlou, A.; Bocchini, P. Computation of bridge seismic fragility by large-scale simulation for probabilistic resilience analysis.
Earthq. Eng. Struct. Dyn. 2015, 44, 1959–1978. [CrossRef]
19. Mosleh, A.; Jara, J.; Razzaghi, M.S.; Varum, H. Probabilistic seismic performance analysis of RC bridges. J. Earthq. Eng. 2020, 24,
1704–1728. [CrossRef]
20. Turmo, J.; Ramos, G.; Aparicio, A.C. Shear truss analogy for concrete members of solid and hollow circular cross section. Eng.
Struct. 2009, 31, 455–465. Available online: http://www.sciencedirect.com/science/article/B6V2Y-4TMYJYT-1/2/8f69b9f83b0
bb052df05ff98d8a38d13 (accessed on 20 December 2022). [CrossRef]
21. Kappos, A.J.; Saiid, S.M.; Nuray, A.M.; Isaković, I. (Eds.) Seismic Design and Assessment of Bridges; Spinger: Berlin/Heidelberg,
Germany, 2012.
22. Deco, A.; Frangopol, D.M. Risk assessment of highway bridges under multiple hazards. J. Risk Res. 2011, 14, 1057–1089. [CrossRef]
23. Dong, Y.; Frangopol, D.M.; Saydam, D. Time-variant sustainability assessment of seismically vulnerable bridges subjected to
multiple hazards. Earthq. Eng. Struct. Dyn. 2013, 42, 1451–1467. [CrossRef]
24. Kim, H.; Sim, S.-H.; Lee, J.; Lee, Y.-J.; Kim, J.-M. Flood fragility analysis for bridges with multiple failure modes. Adv. Mech. Eng.
2017, 9, 1687814017696415. [CrossRef]
25. Ahamed, T.; Duan, J.G.; Jo, H. Flood-fragility analysis of instream bridges—Consideration of flow hydraulics, geotechnical
uncertainties, and variable scour depth. Struct. Infrastruct. Eng. 2021, 17, 1494–1507. [CrossRef]
26. Argyroudis, S.A.; Mitoulis, S.A. Vulnerability of bridges to individual and multiple hazards—Floods and earthquakes. Reliab.
Eng. Syst. Saf. 2021, 210, 107564. [CrossRef]
27. Tanasić, N.; Ilić, V.; Hajdin, R. Vulnerability assessment of bridges exposed to scour. Transp. Res. Rec. 2013, 2360, 36–44. [CrossRef]
28. Anisha, A.; Jacob, A.; Davis, R.; Mangalathu, S. Fragility functions for highway RC bridge under various flood scenarios. Eng.
Struct. 2022, 260, 114244. [CrossRef]
29. George, J.; Menon, A. Analytical fragility curves for displacement-based scour assessment of masonry arch bridges. In Structures;
Elsevier: Amsterdam, The Netherlands, 2022; Volume 46, pp. 172–185.
30. Homaei, F.; Najafzadeh, M. Failure analysis of scouring at pile groups exposed to steady-state flow: On the assessment of
reliability-based probabilistic methodology. Ocean Eng. 2022, 266, 112707. [CrossRef]
31. Homaei, F.; Najafzadeh, M. A reliability-based probabilistic evaluation of the wave-induced scour depth around marine structure
piles. Ocean Eng. 2020, 196, 106818. [CrossRef]
32. Kwon, O.-S.; Elnashai, A. The effect of material and ground motion uncertainty on the seismic vulnerability curves of RC structure.
Eng. Struct. 2006, 28, 289–303. [CrossRef]
33. Bartlett, F.M.; MacGregor, J.G. Statistical analysis of the compressive strength of concrete in structures. Mater. J. 1996, 93, 158–168.
34. Kolisko, J.; Hunka, P.; Jung, K. A statistical analysis of the modulus of elasticity and compressive strength of concrete C45/55 for
pre-stressed precast beams. J. Civ. Eng. Archit. 2012, 6, 1571.
35. Yáñez-Godoy, H.; Mokeddem, A.; Elachachi, S.M. Influence of spatial variability of soil friction angle on sheet pile walls’ structural
behaviour. Georisk Assess. Manag. Risk Eng. Syst. Geohazards 2017, 11, 299–314. [CrossRef]
36. Breysse, D.; Niandou, H.; Elachachi, S.; Houy, L. A generic approach to soil–structure interaction considering the effects of soil
heterogeneity. Geotechnique 2005, 55, 143–150. [CrossRef]
37. Tubaldi, E.; White, C.J.; Patelli, E.; Mitoulis, S.A.; de Almeida, G.; Brown, J.; Cranston, M.; Hardman, M.; Koursari, E.; Lamb, R.;
et al. Invited perspectives: Challenges and future directions in improving bridge flood resilience. Nat. Hazards Earth Syst. Sci.
2021, 22, 795–812. [CrossRef]
38. Diehl, T.H. Potential Drift Accumulation at Bridges; US Department of Transportation, Federal Highway Administration, Research
and Development, Turner-Fairbank Highway Research Center: McLean, VA, USA, 1997.
Water 2023, 15, 129 29 of 29

39. European Standard EN 1993-1-1:2005; Eurocode 3: Design of Steel Structures—Part 1-1: General Rules and Rules for Buildings.
European Committee for Standardization (CEN): Brussels, Switzerland, 2005.
40. European standard EN 1992-1-1:2004; Eurocode 2: Design of Concrete Structures—Part 1-1: General Rules and Rules for Buildings.
European Committee for Standardization (CEN): Brussels, Switzerland, 2004.
41. Herschy, R.W. Streamflow Measurement; CRC Press: Boca Raton, FL, USA, 2008.
42. Zhu, M. The OpenSeesPy Library, Version 3.2.2—2020. 2020. Available online: https://openseespydoc.readthedocs.io/en/latest/
(accessed on 1 November 2022).
43. McKenna, F.; Fenves, G.L.; Scott, M.H. Open System for Earthquake Engineering Simulation (OpenSees); Pacific Earthquake Engineer-
ing Research Center, University of California: Berkeley, CA, USA, 2020.
44. Pais, A.; Kausel, E. Approximate formulas for dynamic stiffnesses of rigid foundations. Soil Dyn. Earthq. Eng. 1988, 7, 213–227.
[CrossRef]
45. Australian Standard AS-5100.2-2004; Bridge Desing—Part 2: Design Loads. Standards Australia: Sydney, Australia, 2004.
46. Yung-Yen, K.; Jiunn-Shyang, C.; Yu-Ching, T.; Cheng-Hsing, C.; Helsin, W.; Chung-Yue, W. Evaluation of Flood-Resistant Capacity
of Scoured Bridges. J. Perform. Constr. Facil. 2014, 28, 61–75. [CrossRef]
47. Hung, C.-C.; Yau, W.-G. Vulnerability evaluation of scoured bridges under floods. Eng. Struct. 2017, 132, 288–299. [CrossRef]
48. Panici, D.; de Almeida, G.A.M. Formation, growth, and failure of debris jams at bridge piers. Water Resour. Res. 2018, 54,
6226–6241. [CrossRef]
49. Hess, J.M. Distribution and Residence Times of Large Woody Debris along South River, Shenandoah Valley, Virginia; University of
Delaware: Newark, DE, USA, 2007.
50. Gurnell, A.; Bertoldi, W. Wood in fluvial systems. In Treatise on Geomorphology; Elsevier: Amsterdam, The Netherlands, 2020.
51. Schwarz, U. Sava white book. In River Sava: Threats and Restoration Potential; EuroNatur/Riverwatch: Radolfzell/Wien, Ger-
many/Austria, 2016.
52. Arneson, L.A.; Zevenbergen, L.W.; Lagasse, P.F.; Clopper, P.E. Evaluating Scour at Bridges. HEC-18, 5th ed.; Hydraulic Engineering
Circular No. 18.; National Highway Institute: Wahington, DC, USA, 2012.
53. Lagasse, P.F.; Zevenbergen, L.W.; Clopper, P.E. Effects of debris on bridge pier scour. In Proceedings of the 5th International
Conference on Scour and Erosion (ICSE-5), San Francisco, CA, USA, 7–11 November 2010; pp. 854–863.
54. Zhang, Y. Probabilistic Structural Seismic Performance Assessment Methodology and Application to an Actual Bridge-Foundation-
Ground System. Ph. D. Thesis, University of California, San Diego, CA, USA, 2006.
55. Melchers, R.E. Structural Reliability Analysis and Prediction; Wiley: Hoboken, NJ, USA, 1999.
56. JCSS. Probabilistic Model Code; Part 3, Resistance models, Section 3.0, Static Properties of Structural Steel; JCSS: 2001. Available
online: https://www.jcss-lc.org/publications/jcsspmc/part_iii.pdf (accessed on 5 December 2022).
57. Aygün, B.; Dueñas-Osorio, L.; Padgett, J.E.; DesRoches, R. Efficient Longitudinal Seismic Fragility Assessment of a Multispan
Continuous Steel Bridge on Liquefiable Soils. J. Bridg. Eng. 2011, 16, 93–107. [CrossRef]
58. Johnson, P.A.; Clopper, P.; Zevenbergen, L.; Lagasse, P.F. Quantifying Uncertainty and Reliability in Bridge Scour Estimations. J.
Hydraul. Eng. 2015, 141. [CrossRef]
59. Vořechovský, M.; Novák, D. Correlation control in small-sample Monte Carlo type simulations I: A simulated annealing approach.
Probabilistic Eng. Mech. 2009, 24, 452–462. [CrossRef]
60. Dolsek, M. Incremental dynamic analysis with consideration of modeling uncertainties. Earthq. Eng. Struct. Dyn. 2009, 38,
805–825. [CrossRef]
61. Dolšek, M. Simplified method for seismic risk assessment of buildings with consideration of aleatory and epistemic uncertainty.
Struct. Infrastruct. Eng. 2012, 8, 939–953. [CrossRef]
62. Kosič, M.; Dolšek, M.; Fajfar, P. Dispersions for the pushover-based risk assessment of reinforced concrete frames and cantilever
walls. Earthq. Eng. Struct. Dyn. 2016, 45, 2163–2183. [CrossRef]
63. Meyerhof, G.G. Some Recent Research on the Bearing Capacity of Foundations. Can. Geotech. J. 1963, 1, 16–26. [CrossRef]
64. Ligthart, F. Failure Mechanisms of Bridge Structures under Natural Hazards; RMIT University: Melbourne, Australia, 2015.
65. Tanasic, N.; Hajdin, R. Performance indicators for bridges exposed to a flooding hazard. In Proceedings of the Joint COST TU
1402—COST TU 1406-IABSE WC1 Workshop, Zagreb, Croatia, 2–3 March 2017.

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy