You are on page 1of 60

Journal Pre-proof

Scour modeling based on immersed boundary method: A pathway to


practical use of three-dimensional scour models

Yalan Song, Yuncheng Xu, Hassan Ismail, Xiaofeng Liu

PII: S0378-3839(21)00180-0
DOI: https://doi.org/10.1016/j.coastaleng.2021.104037
Reference: CENG 104037

To appear in: Coastal Engineering

Received date : 18 May 2021


Revised date : 3 October 2021
Accepted date : 19 October 2021

Please cite this article as: Y. Song, Y. Xu, H. Ismail et al., Scour modeling based on immersed
boundary method: A pathway to practical use of three-dimensional scour models. Coastal
Engineering (2021), doi: https://doi.org/10.1016/j.coastaleng.2021.104037.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the
addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive
version of record. This version will undergo additional copyediting, typesetting and review before it
is published in its final form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.

© 2021 Published by Elsevier B.V.


Journal Pre-proof

of
Scour modeling based on immersed boundary method:

pro
a pathway to practical use of three-dimensional scour
models
Yalan Songa , Yuncheng Xub , Hassan Ismailc , Xiaofeng Liud
a
Department of Civil and Environmental Engineering, Pennsylvania State University, 406 Sackett
Building, University Park, State College, 16802, PA, United States
b
College of Water Resources and Civil Engineering, China Agricultural
University, Beijing, 100083, China
c
d
re-
Seamon Whiteside and Associates, 508 Rhett Street, Greenville, 29601, United States
Department of Civil and Environmental Engineering, Institute of Computational and Data
Science, Pennsylvania State University, 223B Sackett Building, University Park, State
College, 16802, PA, United States

Abstract
lP
Local scour around structures is an important hydraulic phenomenon that in-
volves multiple coupled physical processes, including turbulent flow, sediment
transport, and the evolution of erodible bed. This paper presents a new three-
dimensional (3D) scour model, ibScourFoam, which is based on an immersed
boundary method. This model is capable of dealing with scour around complex
rna

structures and thus it paves the way for practical use of 3D scour models. A spe-
cial wall function has been developed to overcome the non-smoothness problem
of wall shear stress in previous immersed boundary methods. Smooth and accu-
rate wall shear is of critical importance for successful simulation of scour. The
model was validated against cases reported in the literature and one flume ex-
periment conducted in this work. The validation case of flow and scour around
Jou

a vertical pile was reproduced well by our model. The flume experiment is for
scour around a partially-buried, horizontal cylinder placed on a foundation. The

Preprint submitted to Coastal Engineering October 3, 2021


Journal Pre-proof

of
horizontal cylinder is of finite length and the scour is initiated by flow contraction.

pro
The model revealed the distinct stages of scour in the experiment. The complex
process of undermining and exposure of initially buried cylinder and foundation
was captured well by the model.
Keywords:
Hydrodynamics, wall shear stress, sediment transport, scour, complex structures

1. Introduction
1

3
re-
Local scour, induced by accelerated flow field and sediment transport around
coastal and riverine structures, is a common phenomenon with great importance
4 for engineering practice. For example, surveys after major Tsunami and Hur-
5 ricanes have revealed that a major cause of structural failure is foundation scour
lP
6 (FEMA, 2011; Bricker et al., 2012; Tonkin et al., 2013). In 2014, U.S. National In-
7 stitute of Standard and Technology (NIST) identified the erosion and scour around
8 foundations as a key factor for infrastructure resilience against extreme coastal in-
9 undation events (Coulbourne et al., 2014). The severity and extent of scour are
usually estimated with formulas in design guidelines such as U.S. Federal High-
rna

10

11 way Administration’s Hydraulic Engineering Circular-18 (Arneson et al., 2012).


12 Most local scour formulas are of empirical nature. The research and practice
13 communities have long desired to have more reliable and deterministic means to
14 predict the hydrodynamics and morphodynamics that govern the scour process.
15 High-fidelity computational modeling is one of the increasingly popular means
16 thanks to the advancements in computing techniques and power. This paper de-
Jou

17 scribes and demonstrates one such model which has the potential to handle scour
18 around complex structures, a difficult task for many existing models.

2
Journal Pre-proof

of
19 Since the 1970s, many numerical scour and sediment transport models have

pro
20 been developed. They can be classified into one-dimensional (1D) models, two-
21 dimensional (2D) models, and three-dimensional (3D) models. One-dimensional
22 models are not adequate for the simulation of local scour. Some 2D scour models
23 solve 2D Reynolds-Averaged Navier-Stokes equations (RANS) for 2D scour prob-
24 lems, such as scour under pipelines or scour due to submerged wall jets (Liang
25 et al., 2005; Fuhrman et al., 2014; Larsen et al., 2016; Yan et al., 2020). The other
26

27

28
re-
2D scour models solve the depth-averaged shallow water equations (Liu et al.,
2008, 2016). However, vertical fluid motion is very prominent in the vicinity of
structures and is a primary contributor to local scour. Thus, depth-averaged mod-
29 els should theoretically not be used for local scour modeling. In practice, they
30 are often used anyway to obtain flow information, such as velocity magnitude and
lP
31 water depth, to be plugged into empirical formulas to estimate scour depth.
32 Due to the three-dimensional nature of local scour, both hydrodynamically and
33 morphodynamically (Fredsoe and Sumer, 2002), 3D models are naturally needed.
34 A typical 3D scour model consists of at least three parts, namely a hydrodynamic
35 solver to obtain the turbulent flow field of waves and currents, a sediment trans-
rna

36 port solver to calculate sediment fluxes, and a bed evolution solver to update the
37 bed position due to sediment motion (Figure 1). For practical purposes, the hy-
38 drodynamic solver usually uses the Reynolds-Averaged Navier-Stokes equations
39 (RANS) approach to model turbulent flow, although some models also adopt com-
40 putationally more expensive approaches such as Large Eddy Simulation (LES).
41 The sediment transport solver calculates the bed-load transport rate with empir-
Jou

42 ical formulas and the suspended load transport rate with an advection-diffusion
43 equation. The bed evolution part solves the bed sediment mass balance equation,

3
Journal Pre-proof

of
44 i.e., the Exner equation, usually in conjunction with a sand slide treatment to en-

pro
45 sure the resulted bed slope does not exceed the angle of repose (Roulund et al.,
46 2005; Liang et al., 2005; Khosronejad et al., 2011; Jacobsen, 2015; Song et al.,
47 2020b).

re-
lP
Figure 1: Scheme diagram of a scour model: qb is the bed-load transport rate, D
is the deposition rate, and E is the entrainment rate.

48 Computationally, one key difficulty in all 3D scour models is how to track


49 the interface between fluid and sediment. Several approaches have been pro-
rna

50 posed in the literature. One approach is the Arbitrary Lagrangian-Eulerian (ALE)


51 method (Chen, 2006; Liu and Garcı́a, 2008; Liu et al., 2016; Song et al., 2020b).
52 In this approach, the bed surface is a boundary of the computational domain for
53 the flow field. As the bed evolves, the boundary and computational mesh deform
54 accordingly. Though simple, this approach is very limited for practical use. The
55 deformation of mesh may cause excessive stretching and compression such that
Jou

56 mesh quality deteriorates and simulation fails. In addition, the ALE approach can
57 only be used for very simple geometry, such as a vertical pile (Figure 2a) (Roulund
58 et al., 2005; Liu and Garcı́a, 2008; Zhao et al., 2010; Baykal et al., 2017). Fig-

4
Journal Pre-proof

of
59 ure 2b shows the difficulty of the ALE approach for simulating a typical pier

pro
60 foundation with a cap and sub-piles. During a storm event, the foundation may
61 be exposed or re-buried due to scour and sedimentation. Complicated geometric
62 operations have to be performed as the bed boundary sweeps through the foun-
63 dation. It is computationally challenging to use this approach to track the eroded
64 bed and its interaction with the structure.

re-
lP
(a) cylindrical pier (b) Complex pier with a foundation
rna

Figure 2: Schematic diagram of the ALE approach for local scour: (a) simple
vertical pier, (b) complex pier with a cap and sub-piles.

65 The second approach to bed tracking is the two-phase Eulerian-Eulerian model (Link
66 et al., 2012; Nagel et al., 2020). Both water and sediment are modeled as con-
67 tinuous phases, whose interactions can be added into their respective governing
equations. One example is the sedFoam model (Cheng et al., 2017). It is built
Jou

68

69 with the open-source computational physics platform OpenFOAM (OpenFOAM


70 Foundation, 2018). With this approach, the bed interface is implicitly modeled

5
Journal Pre-proof

of
71 and its interaction with structures can be captured without complicated geomet-

pro
72 ric operations. This method has shown good performance in simulating scour
73 processes (Cheng et al., 2017; Chauchat et al., 2017; Nagel et al., 2020). How-
74 ever, this approach heavily relies on the parameterization of non-Newtonian fluid
75 behavior of sediment-water mixture, a field still in need of more research.
76 The third approach to bed tracking is the immersed boundary (IB) method,
77 which is used in this work. The IB method was first proposed in Peskin (1973)
78

79

80
re-
to simulate blood flow in hearts and has flourished since then in many fields of
fluid dynamics. The basic idea of the IB method is to use a fixed background
mesh which does not conform to a solid boundary. Instead, the boundary is “im-
81 mersed” within the simulation domain and its effect on the flow is enforced im-
82 plicitly (Tamaki et al., 2017; Capizzano, 2011; Zhou, 2017; Song et al., 2020a;
lP
83 Xu and Liu, 2021). In the context of scour modeling, the bed is treated as an
84 immersed boundary and can move as scour hole develops. The bed surface is in-
85 dependent of the background mesh and thus the moving bed does not cause any
86 geometric change.
87 The 3D scour model presented in this paper adopts a sharp-interface immersed
rna

88 boundary method. The bed is represented by an unstructured triangular surface


89 mesh embedded in the background mesh (Afzal et al., 2015, 2020; Ahmad et al.,
90 2020; Khosronejad et al., 2011; Song et al., 2020a). Sediment fluxes are calcu-
91 lated on the immersed bed surface based on simulated wall shear stress. Thus,
92 it is critical to obtain accurate and smooth wall shear from the IB method. Non-
93 smoothness is often observed in simulated wall shear due to the arbitrary wall dis-
Jou

94 tance caused by the moving boundary and the different behaviors of wall functions
95 in different sublayers of the near-wall region. Recently, two improved methods for

6
Journal Pre-proof

of
96 smooth wall shear were developed in Song et al. (2020a) and Xu and Liu (2021)

pro
97 for k −  turbulence model. A similar method of Song et al. (2020a) was used in
98 the current 3D scour model under the framework of k − ω model. Additionally,
99 a physically-based slope-limited diffusive sand slide model is used to correct the
100 bed in areas where the slope is larger than the angle of repose (Song et al., 2020b).
101 In this paper, we firstly validated our model by simulating the flow and scour
102 around a vertical pile mounted on the bed. Then, we simulated the scour around
103

104

105
re-
a horizontal cylinder with a complex foundation. In the real world, the complex
foundations are very common, for example large wood debris, complex bridge
piers, and jacket structures in coastal areas. We show the model results match
106 well with experiments. In the second case, we analyzed the undermining process
107 around the foundation and how the exposure of foundation, in turns, impacts the
lP
108 scour process.
109 The organization of this paper is as follows. First we describe the numerical
110 methods of the 3D scour model, including the k − ω SST-SAS turbulence model,
111 the immersed boundary method, the bed morphodynamic solver, and the coupling
112 between the hydrodynamic solver and morphodynamic solver. Then, the scour
rna

113 model is validated against experimental data, followed by an application to scour


114 around a relatively complex structure. This paper concludes with a discussion of
115 the results and future work.

116 2. Numerical Model

The scour model, ibScourFoam, proposed in this work is implemented in


Jou

117

118 OpenFOAM version 5.x (OpenFOAM Foundation, 2018). The immersed bound-
119 ary method part is based on the algorithm in Xu and Liu (2021), which utilizes

7
Journal Pre-proof

of
120 some data structures and code functionalities in the original immersed boundary

pro
121 method in the foam-extend project, a branch of OpenFOAM development (Jasak
122 et al., 2014). The source code and the simulation cases used in this paper are
123 publicly available on GitHub (https://github.com/psu-efd/ibScourFoam).

124 2.1. Hydrodynamic model


125 This section first describes the 3D RANS equations and the k − ω SST-SAS
126 model for turbulence. Then, the immersed boundary method and the near-wall
127

128
re-
treatment are introduced.

2.1.1. Governing equations


129 The 3D incompressible Reynolds-Averaged Navier-Stokes equations can be
130 written as:
lP
∇·u=0 (1)

131

∂u ∇p
+ ∇ · (uu) = − + ∇ · [(ν + νt ) ∇u] (2)
∂t ρ
132 where u is the Reynolds-averaged flow velocity, p is the mean pressure, ρ is water
rna

133 density, ν is kinematic viscosity, and νt denotes the eddy viscosity.


134 The k − ω model has been proven to have better performance in the near-wall
135 region and numerical stability (Wilcox, 2006). In this work, the k − ω SST-SAS
136 model (Egorov and Menter, 2008) is adopted and its governing equations are:
∂k
+ ∇ · (uk) = ∇ · [(ν + αk νt ) ∇k] + Pk − Cµ kω (3)
∂t
137
Jou

∂ω ω
+ ∇ · (uω) =∇ · [(ν + αω νt ) ∇ω] + α Pk − βω2
∂t k (4)
2 1
+ (1 − F1 ) ∇k∇ω + QS AS
αω2 ω

8
Journal Pre-proof

of
138 where k denotes the turbulence kinetic energy, ω denotes the specific dissipation
p

pro
139 rate. Pk = νt S2 is the production term, S = 2Si j Si j is a scalar invariant of the
140 strain rate tensor Sij = 0.5[(∇u)i j + (∇u)i j ]. To reduce the length of the main text,
141 additional terms and coefficients in the model can be found in Appendix A.

142 2.1.2. Immersed boundary method with smooth wall shear


143 The hydrodynamic solver utilizes a background mesh that conforms to the
simulation domain including in-stream structures. However, the mobile sediment
144

145

146

147
re-
bed is modeled as an immersed boundary in the format of an STL (STereoLithog-
raphy) surface (Figure 3). The size of the background mesh needs to be larger than
the physical domain, especially in the vertical direction. This is to leave enough
148 room for the bed to evolve.
lP
rna

Figure 3: Schematic view of a 3D background mesh and an embedded immersed


boundary for a mobile sediment bed.
Jou

149 In the adopted IB method, three different types of cell types can be identified
150 depending on their relative position to the immersed bed surface (see Figure 4): 1)
151 the cells that intersect with the immersed boundary and have their centers located

9
Journal Pre-proof

of
152 on the fluid side are identified as IB cells; 2) the cells fully on the fluid side are

pro
153 identified as fluid cells; and 3) the cells centered on the solid (sediment) side are
154 called solid cells. In the hydrodynamic solver, only the flow variables of fluid
155 cells are solved. Flow variables of solid cells are set to zero. For the IB cells in
156 between, the flow variables are set according to a wall model.
157 In the wall model, three different point types are specified in the near-wall
158 region as shown in Figure 4. The first point is the IB cell center; the second is the
159

160

161
re-
image point (IP) on the normal line to the immersed boundary that passes through
the IB cell center; the third is the hit point (HP), which is the intersection of the
normal line and the immersed boundary. The flow values at IPs are interpolated
162 from their neighboring cells.
163 The distance between an IP and the immersed boundary is set to dIP = 3 l∗ ,
lP
164 where l∗ is the minimum dimension of the IB cell. Through extensive experimen-
165 tation, this distance gives the most reasonable result for all test cases. Near the
166 bed, if the cells have comparable size, this distance is almost uniform. In many
167 existing methods, dIP is set to be proportional to yIB , which is the distance be-
168 tween an IB cell and the immersed boundary. However, since IB cells can have
rna

169 arbitrary distances to the immersed boundary and the distances change when the
170 bed is evolving, dIP in many existing methods can have arbitrary and non-uniform
171 values. One consequence is spiky wall shear stress. As will be shown next, dIP is
172 a key variable in the wall function which controls the accuracy and smoothness of
173 wall shear.
Jou

10
Journal Pre-proof

of
pro
174
re-
Figure 4: Schematic of cell classification in the IB method.

The wall function in the IB method is as follows. It is assumed that an IP and


175 its corresponding IB cell’s center are on the same log-law velocity profile, which
can be written as:
lP
176

1
u+ = ln(E 0 y+ ) (5)
κ
177 where u+ = u/uτ and y+ = yuτ /ν are the dimensionless velocity and wall distance,
178 respectively. The shear velocity is denoted as uτ and the wall shear is τ = ρu2τ . E 0
179 is a wall roughness parameter related to the roughness of the boundary.
rna

180 With the interpolated velocity at an IP, uIP and the distance dIP , the log-law
181 can be written as:
uIP uIP 1 
= + = ln E 0 y+IP (6)
uτ yIP ν/y κ
182 From this equation, the Newton-Raphson method can be used to solve for y+IP
183 iteratively using the following formula (OpenFOAM Foundation, 2018):
y+IP,i + κyuIP /ν
y+IP,i+1 =   (7)
Jou

1 + ln E 0 y+IP,i
184 Here i is the iteration number. Then, the shear velocity, uτ , can be calculated based
185 on the definition of y+IP .

11
Journal Pre-proof

of
186 Another issue in the IB method is the velocity profile between an IB cell center

pro
187 and the immersed bed boundary. If the distance yIB is too small, it may create
188 numerical instability if a log-law profile is used (even if the IB cell is indeed in
189 the log-law layer). To address this, a linear velocity distribution between the IP
190 and HP is assumed for the purpose of assessing the velocity at IB (see Figure 5).
191 We found this is a good and reasonable approximation to the log-law when yIB is
192 not large (which is the case for immersed boundary methods). The slope of the
193

194

195
re-
linear profile is equal to the slope of the log-law profile at the IP. In essence, it
is a linear extension of the log-law velocity profile to the wall. To maintain the
same value of the shear stress on the immersed boundary (u2τ = (ν + νt )∂u /∂y),
196 the eddy viscosity is held constant between IP and HP. This approach has been
197 previously used in Tamaki et al. (2017) and Song et al. (2020a). It has shown good
lP
198 performance in velocity and wall shear stress predictions. Another advantage of
199 this method is that it avoids the discontinuous behavior of different wall functions
200 used in the laminar sublayer and the log-law layer.
rna
Jou

12
Journal Pre-proof

of
pro
re-
Figure 5: Velocity profile in the boundary layer. The dash line is the original
lP
profile. The solid line is the modified profile.

201 Based on the method described above, the tangential velocity, uIB,tan , at the IB
202 cell center is calculated as:
rna

!
∂u+
uIB,tan = uIP,tan − (yIP − yIB )uτ (8)
∂y+ IP,tan

203 where ∂u+ /∂y+ = 1/(κy+ ). The normal velocity, uIB,norm , satisfies the non-penetration
204 condition and is given as:

dIP
uIB,norm = uIP,norm (9)
dIB
p
Jou

205 The turbulence kinetic energy, k = u2τ / Cµ , is approximately constant in the


206 near-wall region (Kalitzin et al. (2005)). Here, Cµ = 0.09. Since the eddy viscosity
207 νt = k/ω is also approximately constant between an IP and an IB cell center to

13
Journal Pre-proof

of
208 maintain a constant shear stress (Tamaki et al., 2017), ωIB is also a constant in this

pro
209 region such that ωIB at the IB cell center can be calculated as:

v
t  2
q !2
6.0ν  k1/2 
ωIB = ωIP = ω2vis + ω2log = +   (10)
0.075y2IP Cµ1/4 κyIP
210 As a proof of the efficacy of the modified wall function described above, a 2D
211 benchmark case for the turbulence flow over a flat plate is presented in Appendix
B. The simulated wall shear and friction coefficient match well with experimental
212

213

214
data. re-
2.2. Bed morphodynamic model

215 Currently, the morphodynamic model can only simulate the bed-load. In all
216 cases shown in this paper, suspended sediment transport is insignificant. The
lP
217 morphodynamics is solved on a unstructured triangular mesh, which is used as
218 the immersed boundary for the flow solver. The bed-load transport rate for a flat,
219 horizontal bed is calculated by the empirical formula proposed in Engelund and
220 Fredsøe (1976),

rna



 1/2 1/2
q0 18.74(θ − θc )(θ − 0.7θc ),

 if θ > θc
p =
 (11)
Rgdd 


0, if θ < θc

221 where R is the submerged specific gravity (1.65 for quartz), d is the sediment grain
222 size, θ = |τ|/(ρgRd) is the Shields number, and θc is the critical Shields number,
223 which can be calculated from the modified fit to the Shields diagram (Parker et al.,
2003)
Jou

224

−0.6
θc0 = 0.5 × (0.22Re−0.6
p + 0.06 × 10−7.7Re p ) (12)

14
Journal Pre-proof

of
225 To include the effect of local bed slope and local flow direction, the critical

pro
226 Shields number is adjusted by
 s 
 sin 2
φ tan2β cos φ sin β 
θc = θc0 cos β 1 − − 
 (13)
µ2s µs

227 where β is the angle of local bed slope, φ is the angle between the velocity and
228 bed slop, and µ s is the static friction coefficient.
229 The evolution of the bed elevation is then solved from the Exner equation,

230
re- (1 − n)
∂zb
∂t
= −∇ · q0
τ
|τ|
!
(14)

where zb is the bed elevation, and n is the porosity of the bed. The Exner equation
231 only considers the conservation of mass for bed sediment. It does not consider
other factors such as the stability of the granular bed if the angle exceeds the angle
lP
232

233 of repose. To correct that, a sand slide model must be applied. In this work, the
234 sand slide model based on slope-limited diffusion is adopted (Song et al., 2020b).

235 2.3. Hydrodynamics and morphodynamics coupling

The flow, sediment transport, and bed evolution are interdependent. Thus, a
rna

236

237 monolithic (fully-coupled) approach in which all of the governing equations are
238 solved simultaneously would be computationally expensive. Similar to most pre-
239 vious scour modeling work, a sequential coupling method is used (see Figure 6).
240 A fully-developed flow field is pre-simulated with an rigid bed as the initial condi-
241 tion. Then, flow and sediment transport are solved sequentially during each time
step. The wall shear stresses at HPs on the immersed boundary are calculated
Jou

242

243 with the wall model. The HPs are not necessary at the centers of triangles of the
244 immersed bed surface mesh. Thus, an interpolation is used to get the wall shear

15
Journal Pre-proof

of
245 at the triangle centers of the bed surface. From there, the sediment transport rate

pro
246 is calculated, and the Exner equation is solved to get the new bed elevation. Once
247 the bed surface is moved, the whole coupling loop starts again.

re-
lP

Figure 6: Flow chart of the scour model


rna

248 For practical purpose, this sequential coupling approach is valid because sed-
249 iment transport is a very slow process, i.e., the morphodynamic time scale (tm )
250 is much larger than that of hydrodynamics (th ) (Liu and Garcı́a, 2008; Termini,
251 2011; Afzal et al., 2015). In fact, most scour models take advantage of this to
252 accelerate the simulation by using a morphological factor α:

α = th /tm , α << 1 (15)


Jou

253 During each time step, the change of bed elevation is amplified by 1/α. The exact
254 value of α to be used may vary from case to case. A balance between computing

16
Journal Pre-proof

of
255 speed and accuracy should be sought. Different values have been used previously.

pro
256 For example, Ponce et al. (1979) suggested the range of 0.01 ≤ α ≤ 0.05. In
257 the context of immersed boundary method, the numerical stability must also be
258 considered. In one time step, the bed elevation change at any location should not
259 exceed the size of an IB cell. This is to avoid spurious oscillations caused by the
260 direct change of a computational cell from solid to fluid (or vice versa) without
261 first transitioning to an intermediate IB cell (Kim et al., 2001; Lee et al., 2011).

262

263

264
re-
2.4. Code implementation

The scour model was implemented in the OpenFOAM version 5.x (Open-
FOAM Foundation, 2018) based on the algorithm of the pisoFoam solver. The
265 scour model contains three parts: the pre-processing of IB information required
by the IB method, the flow solver of pisoFoam with a new wall model described
lP
266

267 in the previous section, and the morphodynamic calculation. The finite volume
268 method (FVM) was used to solve the governing equations of hydrodynamics and
269 morphodynamics on unstructured meshes. The discretization schemes used for
270 simulations are listed in Table 1.
rna

271 3. Validation and Demonstration Cases

272 3.1. Validation case

273 This section shows a 3D validation case of flow and scour around a vertical
274 circular pile reported in Roulund et al. (2005). The flow and scour parts are val-
275 idated separately in the following. The cases shown in this paper are small-scale
Jou

276 problems. Due to the parallel computing capability, large scale applications are
277 feasible given sufficient computing power.

17
Journal Pre-proof

of
Table 1: Summary of discretisation schemes used for simulations (see Open-

pro
FOAM user manual for more details).

Discretisation scheme

Time scheme backward


Gradient scheme Gauss linear
U: Gauss linearUpwind grad(U)

re-
Divergence schemes

Laplacian scheme
k: Gauss upwind
ω: Gauss upwind
Gauss linear corrected
Interpolation scheme linear
Surface normal gradient scheme corrected
lP
278 3.1.1. Flow around a vertical pile mounted on a rigid bed
279 A vertical circular pile is placed on a rigid and flat bed in steady flow as shown
280 in Figure 7. The bed is modeled as an immersed boundary. The validation is
281 mainly focused on whether the model can capture the main flow features which
rna

282 are responsible for scour. The main features are the horseshoe vortex upstream of
283 the pile and the lee-wake vortex downstream. Two cases were validated. The first
284 is with a smooth bed and the second is with a rough bed. Experimental conditions
285 are listed in Table 2.
Jou

18
Journal Pre-proof

of
pro
re-
Figure 7: Schematic view of flow around a vertical circular pile.

Table 2: Simulation conditions for flow around a vertical circular pile on a rigid
lP
bed reported in Roulund et al. (2005).

Parameters Values

Water depth H (cm) 54


Mean flow velocity U (cm/s) 32.6
rna

Pile diameter D (cm) 53.6


Roughness height k s (cm) 0, 1.0
Hydrodynamic time step ∆th (s) 0.005

286 Figure 8 shows the mesh used. The mesh conforms to the pile, i.e., the pile is
part of the background mesh boundary. On the other hand, the bed is modeled as
Jou

287

288 an immersed boundary. This arrangement is optimal because the pile is stationary
289 thus not necessary to be modeled as an immersed boundary which is computa-

19
Journal Pre-proof

of
290 tionally more expensive. The length, width, and height of the domain are 9D, 5D,

pro
291 and 1D, respectively. Here, D is the pile diameter. The mesh has about 7 million
292 cells and is refined near the cylinder, bed, and in the downstream region to better
293 capture the flow structure. The smallest cell size is approximately 1 mm near the
294 cylinder and bed. The largest cell is approximately 50 mm in size near the inlet
295 and outlet. The maximum dimensionless wall distance for the IB cells, y+IBmax , is
296 approximately 13 and 30 for the smooth and rough bed cases, respectively.

re-
lP
Figure 8: Schematic view of mesh used for flow around a vertical circular pile on
a rigid bed.
rna

297 In addition to the immersed bed, the boundary conditions are as follows for
298 the rest of the domain, which include inlet, outlet, two sides of the channel, top
299 surface, and the pile surface:

300 1. At the inlet, profiles for velocity, k, and ω were mapped from a pre-simulated
301 channel flow with no pile.
Jou

302 2. At the outlet, the Neumann boundary conditions were used for all flow vari-
303 ables except the pressure, which was set to zero.
304 3. The two sides were treated as symmetry boundaries.

20
Journal Pre-proof

of
305 4. The shear-free rigid-lid approximation was used for the free surface.

pro
306 5. The pile surface was modeled as a smooth, rigid wall.

307 Figures 9 and 10 show the streamwise velocity at different elevations in the
308 plane of symmetry. The red dots and green dashed lines in the figures represent
309 the experimental measurements and numerical results, respectively, from Roulund
310 et al. (2005). The blue lines represent the results of this work. All the simulations
311 ran for at least 300 s. This is equivalent to approximately 20 flow through time
312

313

314
re-
which is defined as the domain length divided by the average velocity. This en-
sured the simulations reached steady state. Then, the simulations were ran for
more than 100 s (about 7 flow through time) to calculate the mean flow variables.
315 The horseshoe vortex manifests itself as the negative value of the streamwise ve-
locity (pointed to the upstream direction) near the bed (see Figures 9 and 10). The
lP
316

317 comparison is satisfactory and the immersed boundary method captures main flow
318 dynamics well.
rna
Jou

21
Journal Pre-proof

of
pro
re-
lP
rna

Figure 9: Streamwise veloicty, u x , in the plane of symmetry at different elevations


for the smooth, rigid bed case.
Jou

22
Journal Pre-proof

of
pro
re-
lP
rna
Jou

Figure 10: Streamwise veloicty, u x , in the plane of symmetry at different eleva-


tions for the rough, rigid bed case.
23
Journal Pre-proof

of
319 On the downstream side of the pile, the experimental measurement in Roulund

pro
320 et al. (2005) showed an anticlockwise circulation in the smooth bed case, and a
321 clockwise circulation in the rough bed case. These circulations are relatively weak
322 and are difficult to capture by numerical models. For example, the simulated cir-
323 culation in Roulund et al. (2005) was clockwise for the smooth bed case. In our
324 work, the immersed boundary method correctly captured the circulation directions
325 for both smooth and rough bed cases as shown in Figure 11. In addition, the pre-
326

327

328
re-
diction of the downstream velocity was improved in this work (see Figure 9). One
possible reason is that a more refined mesh was used in this work, in comparison
with only 0.2 million cells used in Roulund et al. (2005).
lP
rna
Jou

24
Journal Pre-proof

of
pro
re- (a)
lP

(b)
rna

Figure 11: Velocity field in the plane of symmetry: (a) Smooth bed (b) Rough
bed.

329 Figure 12 shows the comparison of the bed shear stress amplification τ/τo
330 along the x-axis in front of the pile between simulations and the experiment for the
331 smooth bed case. The negative value is due to the horseshoe vortex which makes
Jou

332 the shear pointing upstream. In general, the comparison is acceptable given the
333 dynamic nature of the horseshoe vortex and uncertainties in measurement data.

25
Journal Pre-proof

of
334 Both simulations in this work and Roulund et al. (2005) under-predicted the mag-

pro
335 nitude of the bed shear stress by about 30%. The discrepancy may result from the
336 implementation of two-equation turbulence model in RANS equations that under-
337 estimates the effect of horseshoe vortex system (Khosronejad et al., 2012). It also
338 may be caused by a relatively low order numerical scheme used for the advection
339 terms.

re-
lP

Figure 12: Bed shear stress amplification along the symmetry line upstream of the
pile
rna

340 3.1.2. Scour around a vertical circular pile


341 In this section, the scour model was validated with the experimental case
342 where a pile is mounted on erodible sediment bed (Roulund et al., 2005). The
343 flow resistance estimator in Chen et al. (2019) was used to estimate the roughness
344 height based on the shear velocity measured in the experiment. The diffusivity
Jou

345 for the sand slide model used the same value as in Song et al. (2020b). The mor-
346 phological factor α was adjusted during the simulation because the scour process
347 happens much faster during the early stage than the later stage. Thus, α had a

26
Journal Pre-proof

of
348 value 0.2 at the beginning and then was reduced to 0.05 to accelerate the simula-

pro
349 tion. The simulation ran for 45 mins of real time and took approximately 330 h
350 (14 days) on 80 processors Intel(R) Xeon(R) CPU E5-2680 v3 @ 2.50GHz. The
351 flow condition and sediment properties are shown in Table 3.

Table 3: Parameters for the scour around a vertical circular pile case.

Parameters Values

re-
Water depth H (cm)
Mean flow velocity U (cm/s)
Pile diameter D (cm)
40
46
10
Shear velocity uτ (cm/s) 2.8
Roughness height k s (cm) 0.5
lP
Grain size d50 (mm) 0.26
Critical Shields number for flat bed 0.022
Static friction coefficient µ s 0.63
2
Diffusivity for sand slide (m /s) 0.0005
rna

Hydrodynamic time step ∆th (s) 0.002


Time scale factor α 0.05 ∼ 0.2

352 Figure 13 shows the domain and mesh used for the scour simulation. The
353 domain’s length, width, and height are 16D, 10D, and 3.3D, respectively. The
354 upper part with a height of 2D is the initial fluid domain, and the lower part with
a height of 1.3D is the initial sediment domain. The mesh has about 10 million
Jou

355

356 cells and is refined near the cylinder, bed, and the downstream part. The smallest
357 cell size is approximately 1 mm in the vicinity of the cylinder and the sediment

27
Journal Pre-proof

of
358 bed. The largest cell is approximately 50 mm in size near the inlet and outlet.

pro
359 The maximum dimensionless wall distance for IB cells, y+IBmax , is approximately
360 130. The immersed bed surface mesh has about 60,000 triangles. The boundary
361 conditions are similar to the flow validation case.

re-
lP
Figure 13: Mesh used for scour around a vertical circular pile.

362 Figure 14 shows the contours of bed elevation at four representative times
during the scour process (t = 10 s, 60 s, 150 s, and 600 s). Animations of the
rna

363

364 scour process can be viewed in the Supplementary Materials. The elevation is
365 made dimensionless with the pile diameter, D. The arrows show the direction of
366 the wall shear stress, as well as the direction of the bed-load flux. The reference
367 value of the arrow is the wall shear stress at the inlet, |τ0 |(= ρu2τ ).
368 The four representative scour stages revealed by our model are as follows.
369 First, the scour occurs on the two sides of the pile due to flow contraction and the
Jou

370 horseshoe vortex. Second, the scour hole starts to expand in the upstream direc-
371 tion where the horseshoe vortex strengthens because water flows into the scour

28
Journal Pre-proof

of
372 hole (Ettema et al., 2011) . During this stage, eroded sediment accumulates at the

pro
373 immediate downstream side of the pile. Third, the scour expands to downstream to
374 form a complete hole around the pile. The sediment accumulation downstream of
375 the pile is further washed away due to vortex shedding in the wake. And lastly, the
376 erosion gradually slows down to equilibrium as the scour hole enlarges. Within
377 the scour hole, sediment still moves. However, sediment cannot escape the hole
378 due to sand slide (Song et al., 2020b).

re-
lP
rna
Jou

29
Journal Pre-proof

of
pro
re-
(a) t = 10 s (b) t = 60 s
lP
rna

(c) t = 150 s (d) t = 600 s

Figure 14: Four representative stages of simulated scour hole development. It


shows the bed elevation made dimensionless with the pile diameter D at (a) 10 s,
(b) 60 s, (c) 150 s, and (d) 600 s. The arrows show the direction of wall shear
stress and the bed-load flux.
Jou

379 Figures 15a and 15b show the time evolution of the scour depth at the upstream
380 and downstream side of the pile, respectively. For comparison, the experimental

30
Journal Pre-proof

of
381 and numerical results from Roulund et al. (2005), as well as the simulation results

pro
382 by sedFoam in Nagel et al. (2020) are also plotted. All the numerical models per-
383 form well at the initial scour stage up to 5 mins. The erosion rate of the two-phase
384 flow simulation is smaller than the other simulations and the experiment from 5 to
385 10 mins. All simulations reach equilibrium at around 10 mins. On the upstream
386 side, the two-phase flow simulation gives the maximum zb = 0.63D, whereas
387 the predicted maximum scour depths in Roulund et al. (2005) and ibS courFoam
388

389

390
re-
are 0.91D and 1.06D, respectively. The measured maximum scour depth from
the experiment is 1.25D. On the downstream side, the two-phase flow shows the
maximum zb = 0.18D. Simulations in Roulund et al. (2005) and ibS courFoam
391 both give the maximum zb = 0.6D. The measured maximum scour depth from
392 the experiment is 0.89D. As a result, even though all the numerical models un-
lP
393 derestimate the maximum scour depths at both upstream and downstream sides,
394 the results in Roulund et al. (2005) and ibS courFoam are similar and close to
395 the measured scour depth. One explanation of the underestimation is that two-
396 equation isotropic eddy viscosity turbulence models used to close RANS equa-
397 tions underpredict the effect of the horseshoe vortex system, the main factor of
rna

398 scour around a vertical pile (Khosronejad et al., 2012). The other reason is that
399 the contribution of suspended load to the equilibrium scour depth is not included
400 in these two models (?). In addition, a perturbation was added in the simulation of
401 Roulund et al. (2005) to trigger the development of ripples. However, the impact
402 of ripples was not included in our simulation. This could be another reason that
403 the predicted scour depth in ibS courFoam is slightly less than that in Roulund
Jou

404 et al. (2005). In general, our model captures the scour depth time history rela-
405 tively well.

31
Journal Pre-proof

of
pro
(a) Upstream

re-
(b) Downstream
lP
Figure 15: Time history of scour depth at (a) the upstream and (b) the downstream
side of pile.

406 3.2. Demonstration case


rna

407 To demonstrate the capabilities of the model in dealing with scour around
408 complex structures, a flume experiment was carried out to obtain detailed flow and
409 scour information for comparison purpose. In the experiment, flow and scour were
410 measured around a horizontal cylinder with finite length placed on a foundation
411 (cradle) as shown in Figure 16a. The foundation was fully buried in the sand
412 initially. The cylinder, measuring 0.514 m long and 0.089 m in diameter, was
Jou

413 placed onto the foundation. The position of the cylinder is fixed by the two crosses
414 of the cradle. To prevent excessive undercutting of the cylinder and to ensure it

32
Journal Pre-proof

of
415 engages with the cradle, the cylinder was partially buried initially to a depth of

pro
416 14.5 mm. It was oriented such that the long axis of the cylinder was perpendicular
417 to the primary flow direction. With the development of scour, the bed gradually
418 lowered. As a result, the scour exposed the initially buried part of the cylinder
419 and the cradle. Figure 16b shows the computational mesh. It also shows that as
420 the scour hole develops, the simulated bed (shown as red solid lines) evolves and
421 interacts with the solid boundaries at x = 0 m, where x is the streamwise position
422

423

424
re-
with x = 0 at the center of the cylinder (shown as red dash lines in Figure 16a).
The mesh marked with blue is the initial fluid and the mesh marked with yellow
is the initial sediment.
lP
rna
Jou

33
Journal Pre-proof

of
pro
(a)

re-
lP
(b)

Figure 16: Demonstration case setup: (a) schematic view of the domain for scour
around a horizontal cylinder placed on a foundation. (b) interaction of the bed
rna

with solid boundaries in the background mesh. The red lines are the simulated
bed at different times on the vertical slice of the domain at x = 0 m.

425 Laboratory experiments were conducted in a re-circulating, tilting flume mea-


426 suring 15.24 m in length, 1.52 m in width, and 0.91 m in depth. The test reach
of the flume, extending from 3.048 m to 7.621 m from the inlet, was covered to a
Jou

427

428 depth of 0.12 m with sand having a median grain size of 0.67 mm. Water circu-
429 lated through the system via a horizontal pump capable of delivering a maximum

34
Journal Pre-proof

of
430 flow rate of 240 l/s. Before the test, the sand bed was planed using a scraper

pro
431 attached to an instrument carriage. Once the bed was prepared, the initial bed sur-
432 face including the cylinder and cradle was measured using a handheld 3D scanner
433 (Artec EVA). The same scanner was used to measure the final, eroded bed surface.
434 The flume was slowly filled to a depth of 0.40 m above the sand bed surface,
435 and the pump was turned on. The targeted flow rate was selected by limiting the
436 flow conditions to remain within the clear water scour regime. This meant that the
437

438

439
re-
free-stream flow conditions did not significantly transport bed sediment, and any
erosion was caused by local accelerations. This threshold was first determined
using an estimated Chezy coefficient, then confirmed through preliminary testing.
440 A description of the methodology can be found in Ismail et al. (2021). The flow
441 condition and sediment properties are summarized in Table 4.
lP
442 Once the flow began, the sand bed started to be eroded near the cylinder and
443 cradle and formed a deposit downstream. The flow continued until the bed reached
444 equilibrium at which time there was no noticeable change in the bed surface.
445 Then, measurements of the velocity field were obtained. These include two sets
446 of velocity data: 1) the free-stream inlet velocity far upstream from the structure,
rna

447 and 2) the velocity field in a streamwise, vertical transect extending from x = [-
448 1.03, 1.03]. The streamwise transect measurements were spaced 0.04 m in the
449 streamwise direction and at five elevations from the initial plane bed (15, 30, 45,
450 60, and 75 mm). Note that the streamwise transect did not include data collection
451 immediately above the cylinder.
452 In our experiment, the upstream of the flume was a transitional section filled
Jou

453 with large cobblestones to ensure a gradual flow adjustment from the lower rigid
454 bed to the higher sand bed. The existing of the cobblestones increases the flow

35
Journal Pre-proof

of
pro
Table 4: Test conditions for scour around a horizontal cylinder placed on a cradle
foundation.

re-
Parameters

Water depth H (cm)


Mean flow velocity U (cm/s)
Values

40
22.2
Cylinder diameter D (cm) 8.9
Cylinder length L (cm) 51.4
lP
Shear velocity uτ (cm/s) 1.5
Roughness height k s (cm) 1
Grain size d50 (mm) 0.67
Critical Shields number for flat bed 0.016
rna

Static friction coefficient µ s 0.63


2
Diffusivity for sand slide (m /s) 0.0005
Hydrodynamic time step ∆th (s) 0.005
Time scale factor α 0.002 ∼ 0.02
Jou

36
Journal Pre-proof

of
455 resistance of the bed, as well as the Nikuradse equivalent roughness at the up-

pro
456 stream region of the sand bed. Thus, the hydraulic roughness is related to not only
457 sediment sizes, but also incoming flow state. In this work, we adapted the value
458 of roughness k s by fitting the inlet velocity profile measured from the experiment,
459 instead of only using the sediment size d50 . Figure 17 shows the mean velocity
460 profiles upstream of the test section. The red dots are the measured velocity. The
461 blue dots are the pre-simulated profile in the channel with no object. The black
462

u+ =
re-
line is the velocity profile fitted with the general log-law

1
κ
u
ln(y+ ) + B − ∆B(k+s ), u+ = + , y+ =
u
uτ y
ν
(16)

463 where B is 5.2. The roughness function ∆B depends on k+s = k s uτ /ν and represents
464 the downward shift of the logarithmic velocity profile due to roughness. The
lP
465 functional form of ∆B in Cebeci and Bradshaw (1977) was used:






0 if k+s < 2.25




∆B = 
[B − 8.5 + 1κ ln k+s ] sin[0.4258(ln k+s − 0.811)] if 2.25 ≤ k+s < 90 (17)







 B − 8.5 + 1 ln k+s if k+s ≥ 90
rna

466 The fitted shear velocity uτ is 0.015 m/s and the Nikuradse equivalent sand rough-
467 ness k s is 0.01 m.
Jou

37
Journal Pre-proof

of
pro
re-
Figure 17: Mean velocity profile measured at the undisturbed inlet of flume
lP
468 In the model, the inlet condition mapped from the uniform-channel flow cal-
469 culation with the fitted roughness k s is a standard logarithmic profile. Even though
470 the measured velocity profile deviates from the log law in the upper region, the
rna

471 effect of different inlet conditions used in the model and the experiment is not
472 significant since the scour is induced by flow contraction near the object.
473 The simulation domain has a length, width, and height of 2.2 m, 1.2 m, and
474 5.2 m, respectively. The mesh has approximately 8 million cells and the smallest
475 cell is approximately 1.6 mm in size. The maximum dimensionless wall distance
476 for IB cells, y+IBmax , is approximately 40. The immersed bed surface has 60,000
triangles. The simulation ran for 17 hours of real time and took approximately
Jou

477

478 170 h (about 7 days) on 200 processors (Intel(R) Xeon(R) CPU E5-2680 v3 @
479 2.50GHz).

38
Journal Pre-proof

of
480 Figure 18 shows four stages of scour from the initial to equilibrium in four

pro
481 rows. The columns, in their order, show the following: slices of bed at x = 0 m
482 (column a), scour depth patterns (column b), streamlines (column c), wall shear
483 stress amplification factors (column d), dimensionless sediment flux (column e),
484 and photos from experiment (column f). Arrows in column d show the direction
485 of wall shear stress. The reference value of the arrow is the wall shear stress at the
486 inlet |τ0 |(= ρu2τ )
487

488

489
re-
With the details revealed by the computational model, the four stages of ero-
sion observed in the laboratory experiment and laid out in Figure 18 can be further
examined:

490 1. Wake flow stage (Figure 18, row 1): water flows over the cylinder and con-
tracts when passing around the two ends of the cylinder. The approaching
lP
491

492 flow decelerates in front of the cylinder. Most of the incoming water flows
493 up and over the cylinder. A small portion of incoming water flow down and
494 forms a weak vortex that erodes the sand at the upstream side the cylinder.
495 In this stage, most sediment movement and erosion happen at the two ends
rna

496 of the cylinder where flow contracts. As sediment moves away, the two ends
497 are undermined and the underside of the cylinder starts to be exposed.
498 2. First jet flow stage (Figure 18, row 2): In this stage, the dominant flow is the
499 two jet flows under the two ends of the cylinder. Once the scour holes there
500 are developed, the incoming flow is split when it encounters the cylinder.
501 The downward flow into the scour holes contracts and accelerates to form
Jou

502 two jet flows under the cylinder. The sediment in the scour holes is pushed
503 out and deposited in the adjacent downstream side. It is also observed that
504 some sediment near the two ends of the cylinder slides down to the deep

39
Journal Pre-proof

of
505 part of scour hole and then is washed away by the jet flow. Gradually, the

pro
506 scour holes on the two ends expand toward the center.
507 3. Second jet flow stage (Figure 18, row 3): The two scour holes merge at
508 the center and expose the two crosses of the cradle. More incoming flow
509 goes under the cylinder and contacts at the center due to the existence of
510 two crosses. At this stage, the dominant flow is the one single jet under the
511 whole cylinder. The sediment flux is maximum and thus the scour rate is
512

513

514
re-
the largest under the center of the cylinder because of the cradle’s blocking
effect. As a result, more sediment is being moved out from the center and
then deposited downstream. The newly deposited sediment merges with
515 the sediment from previous two stages and form a ridge downstream. With
516 time, the bed under the cylinder slightly flattens.
lP
517 4. Pier-like flow stage (Figure 18, row 4): In this stage, undermining under
518 the cylinder continues but mainly happens around the two crosses of the
519 supporting cradle. The two crosses act like bridge piers and two local scour
520 holes around them start to form. The scour pattern around the two crosses
bears some resemblance to the vertical pile scour shown in Figure 14. Scour
rna

521

522 continues near the cradle until equilibrium and the deepest part of the scour
523 hole is around the two crosses. During the process, the sediment that moved
524 out of the center region deposited in the bar downstream and forms the
525 highest point in the middle.
Jou

40
a b c d e f
Stage 1
t = 0.4 hr
Jou
Stage 2
t = 1.4 hr
rna

42
Stage 3
t = 3.6 hr
lP
Stage 4
Journal Pre-proof

t = 17.6 hr
re-
Figure 18: Different stages of the simulated scour process around a horizontal cylinder supported by a cradle
foundation. Rows corresponds to different stages. Columns a, b, c, d, e, and f are bed profile, scour depth contour,
pro
flow streamlines, wall shear amplification, sediment flux, and photos from experiment, respectively.
of
Journal Pre-proof

of
526 It should be noted that the wall shear stress distribution in Figure 18 (column

pro
527 d) is smooth, which enables a successful scour simulation. This demonstrates
528 the capability of the immersed boundary method and the associated wall model
529 proposed in this work.
530 Figure 19 shows a good comparison of the scour profile on the vertical slice
531 at x = 0 m. The maximum scour depth is well captured near the crosses of the
532 cradle. Bed slope of the scour hole is also consistent with the experimental data.
533

534

535
re-
The main discrepancy is that the scour hole on the right side is slightly deeper
than that on left side in the experiment. It may result from the slight asymmetry of
experimental setup. For comparison, the simulated result is symmetric along the
536 symmetry plane of the flume. In this case, the driving flow of scour is dominated
537 by flow contraction, rather than the horseshoe vortex that is underestimated by the
lP
538 RANS model. Thus, a better agreement between the simulation and experiment
539 is produced by our model than that of the previous scour case.
rna
Jou

Figure 19: Equilibrium scour profile on the vertical slice at x = 0 m.

540 We made more comparisons between simulations and experiments. To reduce

43
Journal Pre-proof

of
541 the length, they can be found in Appendix C.

pro
542 4. Limitations

543 Despite the success shown in this paper, the current model still has limitations.
544 The major one is it is computationally more demanding than previous methods,
545 such as the ALE method. In the current model, the simulation domain has to be
546 larger to include the sediment part. In addition, the mesh in the sediment part has
547

548

549
re-
to be refined to have accurate results of flow and wall shear stress. As the Reynolds
number increases, the required mesh resolution of the immersed boundary method
increases much faster than the ALE method (Mittal and Iaccarino, 2005). The
550 other limitation is the numerical instability caused by the switch among the IB,
551 fluid, and solid cells when the boundary moves. Our background mesh is very
lP
552 stretched in the vertical direction to increase the mesh resolution and decrease
553 the total mesh size, which also increases the instability of our model. There-
554 fore, we used a relatively low order numerical scheme for the advection terms
555 in Table 1. This scheme adds more numerical diffusion into the model and de-
creases its capability to capture finer flow features downstream. In addition, scour
rna

556

557 models solving RANS equations with two-equation turbulence models underesti-
558 mate the wall shear, as well as the scour depth induced by the horseshoe vortex
559 system (Khosronejad et al., 2012). To alleviate this problem, a conservative mod-
560 ification of Brownlie’s (?) curve from Parker et al. (2003) was used to calculate
561 the critical shields number, which amplifies the bed-load transport rate and in-
creases the estimated maximum scour depth. The current model also lacks the
Jou

562

563 suspended sediment component. Suspended sediment transport within the frame-
564 work of immersed boundary method, especially how entrainment and deposition

44
Journal Pre-proof

of
565 can be modeled, is currently under development.

pro
566 5. Conclusions

567 A 3D scour model, ibScourFoam, has been developed based on a sharp-interface


568 immersed boundary method. It can simulate the scour process around complex
569 structures, a difficult task for many previous scour models. This paper describes
570 the details of the model and its implementation. In particular, the wall model
571

572

573
scour simulations.
re-
proposed here produces accurate and smooth wall shear stress, a prerequisite for

The model was validated by simulating the flow and scour cases in Roulund
574 et al. (2005). The flow structures and the temporal evolution of scour depth were
575 well predicted. The model was further validated with a more complicated flume
lP
576 experiment case where a short cylinder sits on a cradle. The cylinder is initially
577 partially buried and the cradle is fully buried. The model successfully simulated
578 the undermining process starting at the two ends of the cylinder, the progressive
579 enlargement of the scour holes, the merging of the two scour holes, and the sub-
sequent scour around the fully exposed crosses of the cradle. With the 3D scour
rna

580

581 simulation results, it is now possible to analyze the details of the process and
582 different stages of the scour development.
583 The 3D scour model developed in this work provides a viable path for scour
584 modeling to be used in engineering practice. It helps us go beyond the “scour
585 around simple geometry” phase in this field. With ever increasing computing
power, the relatively high computational cost required by the immersed boundary
Jou

586

587 method will soon become more affordable.

45
Journal Pre-proof

of
588 Appendix A. Terms and coefficients in the turbulence model

pro
589 Additional terms and coefficients in the k − ω SST-SAS model are described
590 below. The blending function, F1 , is calculated as:
   √   
4




   k 500ν  4αω2 k 
 

F1 = tanh 

 min max  ∗ , 2  ,   (A.1)

 β ωy y ω 2αω2 ω ∇k∇ωy  
−1 2  

591 QS AS is the additional SAS source term which has the form of
 !2 ! 

592

where L = k/(c0.25
re- 
QS AS = max ζ2 κS2
L
LνK

2Ck
σΦ
max
|∇ω|2 |∇k|2
ω2
,
k 2

, 0 (A.2)

µ ω) is the modeled length scale. LνK is a three-dimensional

593 generalization of the classic one-dimensional boundary layer definition and is cal-
594 culated as:
p
lP
LνK = max(κS/|∇2 ū|, CS δ ζ2 κ/(β/β∗ − γ)) (A.3)

595 where δ is the cubic root of computational cell volume.


596 The eddy viscosity, νt , is then calculated as:

a1 k
νt = (A.4)
max(a1 ω, b1 SF2 )
rna

597 where the second blending function, F2 , is calculated as:


  √ 2 


  2 k 500ν  


F2 = tanh 
 max  ∗ , 2   (A.5)
 β ωy y ω  

598 The coefficients used in the turbulence formulas are listed as follows: αk = F1 (αk1 −
599 αk2 ) + αk2 , αk1 = 0.85, αk2 = 1.0, Cµ = 0.09, αω = F1 (αω1 − αω2 ) + αω2 , αω1 =
Jou

600 0.5, αω2 = 0.856, α = 0.8, β = F1 (β1 − β2 ) + β2 , β1 = 0.075, β2 = 0.0828, β∗ =


601 0.09, ζ2 = 3.51, κ = 0.41, C = 2, σΦ = 2/3, CS = 0.11, γ = F1 (γ1 − γ2 ) +
602 γ2 , γ1 = 5/9, γ2 = 0.44, a1 = 0.31, and b1 = 1.0.

46
Journal Pre-proof

of
603 Appendix B. Validation of wall function

pro
604 The turbulent flow over a flat plate is a widely used 2D case to validate the wall
605 shear prediction of hydrodynamic solvers. A 2D boundary layer without pressure
606 gradient was tested at ReL = 5 × 106 . The length, L, of the plate is 2 m. A slip
607 surface with a length of 0.1L is added in front of the flat plate to generate a free
608 stream. The upper boundary is 0.4L away from the flat plate. We refined the
609 mesh near the leading edge of the flat plate and the boundary layer region (0.2L).
610

611

612
re-
A schematic view of the simulation mesh is shown in Fig. B.20. The mesh is
aligned with the flat plate. N is the number of cells and R is the cell expansion
ratio (R is the ratio of the size of the end cell ∆N to the size of the start cell ∆1
613 along the edge direction). Five different cell expansion ratios were tested in the
614 y-direction of the boundary layer region to change the size of the first wall cell.
lP
rna

Figure B.20: Schematic view of mesh for flow over a flat plate

615 In the simulation, the slip surface uses symmetric boundary condition and the
Jou

616 flat plate is set as an immersed boundary. The proposed wall model is applied
617 on the flat plate. Figure B.21 shows the predicted skin friction coefficient, C f ,
618 compared with the experimental data from Wieghardt and Tillmann (1951). As

47
Journal Pre-proof

of
619 the y+IB decreases due to the progressive refinement of the mesh, the simulation

pro
620 results match closely to the experimental data. When the averaged y+IB is located
621 in the laminar layer, the prediction of the shear velocity is still good, especially at
622 the leading edge where the flow is dominantly laminar.

re-
Figure B.21: Skin friction coefficient (C f = τw /(1/2ρU 2 )) along the flat plate.
lP
623 Appendix C. More comparison for the cylinder on a cradle case

624 The flow field at scour equilibrium was also measured in the experiment. It is
used here to compare with the simulation results. The velocity under the cylinder
rna

625

626 is unable to be measured by the Nortec Vectrino Plus Acoustic Doppler Velocime-
627 ter (ADV). Figure C.22 shows the streamwise velocity contour on the center slice
628 (y = 0 m) at equilibrium. The tail of the jet flow at the downstream of the cylinder
629 is in relatively good agreement with the experiment. The back flow circulation
630 behind the cylinder is not as strong as that in the experiment. This may be due
to the use of RANS model and the mesh behind the cylinder is not fine enough.
Jou

631

632 Nonetheless, the bed profiles (black lines in Figure C.22) compare well.

48
Journal Pre-proof

of
pro
(a)

re-
lP
(b)

Figure C.22: Contour of streamwise velocity, u x , on the center slice at equilibrium:


(a) Simulation, (b) Experiment. The black lines are the bed profiles.
rna

633 Figure C.23 shows the comparison of the bed contours at equilibrium from the
634 simulation and the experiment. Though not an exact match, it can be observed that
635 the features of the scour hole and deposition bar are well captured by the model.
636 Even the details of the scour holes around the crosses of the cradle are captured.
637 In both the experiment and the simulation, a large deposition bar appears in the
downstream, and the peak bed elevation is at the center. The simulated deposition
Jou

638

639 height behind the cylinder is lower than the experimental value. This could be due
640 to the coarse mesh used in the downstream side of the domain.

49
Journal Pre-proof

of
pro
re-
(a) (b)

Figure C.23: Bed contours at scour equilibrium: (a) Simulation (b) Experiment.

641 Acknowledgements

This work is supported by the Strategic Environmental Research and Devel-


lP
642

643 opment Program (SERDP, Award Number W74RDV70063408).

644 Supplementary Materials

645 Supplementary materials (animation videos) associated with this article can
rna

646 be found in the online version. Source code and cases can be found on GitHub
647 (https://github.com/psu-efd/ibScourFoam).

648 References

649 Afzal, M.S., Bihs, H., Kamath, A., Arntsen, Ø.A., 2015. Three-dimensional nu-
650 merical modeling of pier scour under current and waves using level-set method.
Jou

651 Journal of Offshore Mechanics and Arctic Engineering 137, 032001.

50
Journal Pre-proof

of
652 Afzal, M.S., Bihs, H., Kumar, L., 2020. Computational fluid dynamics modeling

pro
653 of abutment scour under steady current using the level set method. International
654 Journal of Sediment Research 35, 355–364.

655 Ahmad, N., Kamath, A., Bihs, H., 2020. 3D numerical modelling of scour around
656 a jacket structure with dynamic free surface capturing. Ocean Engineering 200,
657 107104.

658

659

660
re-
Arneson, L.A., Zevenbergen, L.W., Lagasse, P.F., Clopper, P.E., 2012. Evaluat-
ing scour at bridges. Technical Report FHWA-HIF-12-003, HEC-18. Federal
Highway Administration, Washington, DC.

661 Baykal, C., Sumer, B.M., Fuhrman, D.R., Jacobsen, N.G., Fredsøe, J., 2017. Nu-
662 merical simulation of scour and backfilling processes around a circular pile in
lP
663 waves. Coastal Engineering 122, 87–107.

664 Bricker, J.D., Francis, M., Nakayama, A., 2012. Scour depths near coastal
665 structures due to the 2011 Tohoku Tsunami. J. Hydraul. Res. 50, 637–641.
666 doi:10.1080/00221686.2012.721015.
rna

667 Capizzano, F., 2011. Turbulent wall model for immersed boundary methods.
668 AIAA journal 49, 2367–2381.

669 Cebeci, T., Bradshaw, P., 1977. Momentum transfer in boundary layers. Wash-
670 ington .

Chauchat, J., Cheng, Z., Nagel, T., Bonamy, C., Hsu, T.J., 2017. SedFoam-2.0:
Jou

671

672 a 3-D two-phase flow numerical model for sediment transport. Geoscientific
673 Model Development 10, 4367–4392.

51
Journal Pre-proof

of
674 Chen, B., 2006. The numerical simulation of local scour in front of a vertical-wall

pro
675 breakwater. Journal of Hydrodynamics 18, 132–136.

676 Chen, Y., DiBiase, R.A., McCarroll, N., Liu, X., 2019. Quantifying flow resis-
677 tance in mountain streams using computational fluid dynamics modeling over
678 structure-from-motion photogrammetry-derived microtopography. Earth Sur-
679 face Processes and Landforms 44, 1973–1987.

680

681

682
re-
Cheng, Z., Hsu, T.J., Calantoni, J., 2017. SedFoam: A multi-dimensional Eulerian
two-phase model for sediment transport and its application to momentary bed
failure. Coastal Engineering 119, 32–50.

683 Coulbourne, W., Galsworthy, J., Hangan, H., Jones, C., Letchford, C., Smith,
684 T., 2014. Measurement Science R&D Roadmap for Windstorm and Coastal
lP
685 Inundation Impact Reduction. Technical Report 14-973-13. National Institute
686 of Standards and Technology, Gaithersburg, MD.

687 Egorov, Y., Menter, F., 2008. Development and application of SST-SAS turbu-
688 lence model in the DESIDER project, in: Advances in Hybrid RANS-LES
rna

689 Modelling. Springer, pp. 261–270.

690 Engelund, F., Fredsøe, J., 1976. A sediment transport model for straight alluvial
691 channels. Nordic Hydrology 7, 293–306.

692 Ettema, R., Melville, B.W., Constantinescu, G., 2011. Evaluation of bridge scour
693 research: Pier scour processes and predictions. Technical Report Project 24-27.
Jou

694 National Cooperative Highway Research Program, Washington, DC.

695 FEMA, 2011. Coastal construction manual: Principles and practices of planning,

52
Journal Pre-proof

of
696 siting, designing, constructing, and maintaining residential buildings in coastal

pro
697 areas. Technical Report P-55. Federal Emergency Management Agency.

698 Fredsoe, J., Sumer, B.M., 2002. The mechanics of scour in the marine environ-
699 ment. volume 17. World Scientific Publishing Company.

700 Fuhrman, D.R., Baykal, C., Sumer, B.M., Jacobsen, N.G., Fredsøe, J., 2014. Nu-
701 merical simulation of wave-induced scour and backfilling processes beneath
submarine pipelines. Coastal engineering 94, 10–22.
702

703

704
re-
Ismail, H., Xu, Y., Liu, X., 2021. Flow and scour around idealized porous engi-
neered log jam structures. Journal of Hydraulic Engineering 147, 04020089.

705 Jacobsen, N.G., 2015. Mass conservation in computational morphodynamics:


uniform sediment and infinite availability. International Journal for Numerical
lP
706

707 Methods in Fluids 78, 233–256.

708 Jasak, H., Rigler, D., Tuković, Ž., 2014. Design and implementation of Immersed
709 Boundary Method with discrete forcing approach for boundary conditions, in:
710 11th World Congr Comput Mech WCCM 2014, 5th Eur Conf Comput Mech
rna

711 ECCM 2014 6th Eur Conf Comput Fluid Dyn ECFD 2014.

712 Kalitzin, G., Medic, G., Iaccarino, G., Durbin, P., 2005. Near-wall behavior of
713 rans turbulence models and implications for wall functions. Journal of Compu-
714 tational Physics 204, 265–291.

715 Khosronejad, A., Kang, S., Borazjani, I., Sotiropoulos, F., 2011. Curvilinear
Jou

716 immersed boundary method for simulating coupled flow and bed morphody-
717 namic interactions due to sediment transport phenomena. Advances in Water
718 Resources 34, 829–843.

53
Journal Pre-proof

of
719 Khosronejad, A., Kang, S., Sotiropoulos, F., 2012. Experimental and compu-

pro
720 tational investigation of local scour around bridge piers. Advances in Water
721 Resources 37, 73–85.

722 Kim, J., Kim, D., Choi, H., 2001. An immersed-boundary finite-volume method
723 for simulations of flow in complex geometries. Journal of computational
724 physics 171, 132–150.

Larsen, B.E., Fuhrman, D.R., Sumer, B.M., 2016. Simulation of wave-plus-


725

726

727
re-
current scour beneath submarine pipelines. Journal of Waterway, Port, Coastal,
and Ocean Engineering 142, 04016003.

728 Lee, J., Kim, J., Choi, H., Yang, K.S., 2011. Sources of spurious force oscillations
729 from an immersed boundary method for moving-body problems. Journal of
lP
730 computational physics 230, 2677–2695.

731 Liang, D., Cheng, L., Li, F., 2005. Numerical modeling of flow and scour below
732 a pipeline in currents: Part ii. scour simulation. Coastal engineering 52, 43–62.

733 Link, O., González, C., Maldonado, M., Escauriaza, C., 2012. Coherent structure
rna

734 dynamics and sediment particle motion around a cylindrical pier in developing
735 scour holes. Acta Geophysica 60, 1689–1719.

736 Liu, M., Lu, L., Teng, B., Zhao, M., Tang, G., 2016. Numerical modeling of
737 local scour and forces for submarine pipeline under surface waves. Coastal
738 Engineering 116, 275–288.
Jou

739 Liu, X., Garcı́a, M.H., 2008. Three-dimensional numerical model with free water
740 surface and mesh deformation for local sediment scour. Journal of Waterway,
741 Port, Coastal, and Ocean Engineering 134, 203–217.

54
Journal Pre-proof

of
742 Liu, X., Landry, B., Garcı́a, M.H., 2008. Two-dimensional scour simulations

pro
743 based on coupled model of shallow water equations and sediment transport on
744 unstructured meshes. Coastal Engineering 55, 800–810.

745 Mittal, R., Iaccarino, G., 2005. Immersed boundary methods. Annu. Rev. Fluid
746 Mech. 37, 239–261.

747 Nagel, T., Chauchat, J., Bonamy, C., Liu, X., Cheng, Z., Hsu, T.J., 2020. Three-
748

749

750
re-
dimensional scour simulations with a two-phase flow model. Advances in Water
Resources 138, 103544.

OpenFOAM Foundation, 2018. OpenFOAM v5 User Guide. URL: https://


751 cfd.direct/openfoam/user-guide.
lP
752 Parker, G., Toro-Escobar, C.M., Ramey, M., Beck, S., 2003. Effect of floodwater
753 extraction on mountain stream morphology. Journal of Hydraulic Engineering
754 129, 885–895.

755 Peskin, C., 1973. Flow patterns around heart valves: a digital computer method for
solving the equations of motion. IEEE transactions on biomedical engineering
rna

756

757 BME-20, 316–317.

758 Ponce, V.M., Simons, D.B., Garica, J.L., 1979. Modeling alluvial channel bed
759 transients. Journal of the Hydraulics Division 105, 245–256.

760 Roulund, A., Sumer, B.M., Fredsøe, J., Michelsen, J., 2005. Numerical and ex-
perimental investigation of flow and scour around a circular pile. Journal of
Jou

761

762 Fluid Mechanics 534, 351–401.

55
Journal Pre-proof

of
763 Song, Y., Lai, Y.G., Liu, X., 2020a. An improved immersed boundary method

pro
764 for simulating flow hydrodynamics in streams with complex terrains. Water 12,
765 2226.

766 Song, Y., Xu, Y., Liu, X., 2020b. Physically based sand slide method in scour
767 models based on slope-limited diffusion. Journal of Hydraulic Engineering
768 146, 04020074.

769

770

771
re-
Tamaki, Y., Harada, M., Imamura, T., 2017. Near-wall modification of spalart–
allmaras turbulence model for immersed boundary method. AIAA Journal 55,
3027–3039.

772 Termini, D., 2011. Bed scouring downstream of hydraulic structures under steady
773 flow conditions: Experimental analysis of space and time scales and implica-
lP
774 tions for mathematical modeling. Catena 84, 125–135.

775 Tonkin, S.P., Francis, M., Bricker, J.D., 2013. Limits on coastal scour depths due
776 to tsunami, in: Int. Efforts Lifeline Earthq. Eng., American Society of Civil
777 Engineers, Reston, VA. pp. 671–678. doi:10.1061/9780784413234.086.
rna

778 Wieghardt, K., Tillmann, W., 1951. On the turbulent friction layer for rising
779 pressure .

780 Wilcox, D.C., 2006. Turbulence modeling for CFD. DCW Industries, La Canada,
781 CA.

Xu, Y., Liu, X., 2021. An immersed boundary method with y+-adaptive wall
Jou

782

783 function for smooth wall shear. International Journal for Numerical Methods in
784 Fluids 93, 1929– 1946.

56
Journal Pre-proof

of
785 Yan, X., Mohammadian, A., Rennie, C.D., 2020. Numerical modeling of local

pro
786 scour due to submerged wall jets using a strict vertex-based, terrain confor-
787 mal, moving-mesh technique in openfoam. International Journal of Sediment
788 Research 35, 237–248.

789 Zhao, M., Cheng, L., Zang, Z., 2010. Experimental and numerical investigation
790 of local scour around a submerged vertical circular cylinder in steady currents.
791 Coastal Engineering 57, 709–721.

792

793
re-
Zhou, C., 2017. RANS simulation of high-re turbulent flows using an immersed
boundary method in conjunction with wall modeling. Computers & Fluids 143,
794 73–89.
lP
rna
Jou

57
Journal Pre-proof

of
Highlights
Scour modeling based on immersed boundary method: a pathway to practical use of three-
dimensional scour models

Yalan Song, Yuncheng Xu, Hassan Ismail, Xiaofeng Liu

pro
 An open-source three-dimensional (3D) scour model based on an immersed boundary
method has been developed.
 The model utilizes a new algorithm to produce smooth and accurate wall shear, which is
key for sediment transport.
 The model can simulate scour around structures with complex geometries under waves
and currents, paving the way for practical use of 3D scour modeling.

re-
lP
rna
Jou
Journal Pre-proof

of
Yalan Song: Software, Investigation, Visualization, Writing
Yuncheng Xu: Software, Writing
Hassan Ismail: Experiment, Writing
Xiaofeng Liu: Conceptualization, Methodology, Supervision

pro
re-
lP
rna
Jou
Journal Pre-proof

of
Declaration of interests

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

pro
☐The authors declare the following financial interests/personal relationships which
may be considered as potential competing interests:

re-
lP
rna
Jou

You might also like