Вы находитесь на странице: 1из 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/270466345

A bridge Pier Scour Model with non-uniform Sediments

Article  in  Water Management · October 2014


DOI: 10.1680/wama.13.00094

CITATION READS

1 373

3 authors:

Shaghayegh Pournazeri Samuel Li


University of Iowa Concordia University Montreal
9 PUBLICATIONS   17 CITATIONS    42 PUBLICATIONS   134 CITATIONS   

SEE PROFILE SEE PROFILE

Fariborz Haghighat
Concordia University Montreal
341 PUBLICATIONS   5,917 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Nanoparticle filtration - Application to collective ventilation View project

Heating Demand Prediction of District Systems View project

All content following this page was uploaded by Shaghayegh Pournazeri on 03 June 2015.

The user has requested enhancement of the downloaded file.


Water Management Proceedings of the Institution of Civil Engineers

A bridge pier scour model with non-uniform http://dx.doi.org/10.1680/wama.13.00094


sediments Paper 1300094
Pournazeri, Li and Haghighat Received 31/08/2013 Accepted 10/02/2014
Keywords: hydraulics & hydrodynamics/mathematical modelling/river
engineering

ICE Publishing: All rights reserved

A bridge pier scour model


with non-uniform sediments
Shaghayegh Pournazeri MASc Fariborz Haghighat PhD, PEng, FASHRAEF, FISIAQ
Graduate Student, Department of Building, Civil and Environmental Professor, Department of Building, Civil and Environmental Engineering,
Engineering, Concordia University, Montreal, Canada Concordia University, Montreal, Canada
S. Samuel Li PhD, PEng, MIAHR, MCSCE
Associate Professor, Department of Building, Civil and Environmental
Engineering, Concordia University, Montreal, Canada

Pier scour is a significant problem affecting the safety of bridges. For given hydraulic and geometric conditions,
accurate prediction of scour with non-uniform sediments is important, but this need has not been fulfilled. The
purpose of this research was to develop a three-dimensional model for scour prediction and to verify the model
using laboratory measurements. The model allows for selective transport of non-uniform sediments, particle
hiding and bed-level change in response to scour and deposition. The development of scouring around a circular
pier on a mobile channel bed with non-uniform sediments was successfully predicted and scour depth prediction
agreed well with the measurements. It was found that scour patterns emerge from the lateral sides of the pier
and migrate towards its upstream nose. Upstream of the pier, strong downflow and vortex motions develop and
effectively remove sediments from the foot of the pier; at equilibrium, the bed-surface slope almost reaches the
angle of repose of sediments. On the upstream side, the scour hole has the shape of almost half a cone. Grain
size non-uniformity reduces the magnitude of scour. These findings are of relevance to the safe and cost-effective
design of pier foundations. The modelling techniques are computationally efficient and are useful for field-scale
application.

Notation u friction velocity (m/s)


D pier diameter (m) up magnitude of (Reynolds-averaged) velocity
dg geometric mean grain size (mm) parallel to the bed surface (m/s)
dj grain size of the jth sediment fraction (mm) uo cross-sectionally averaged approach flow
ds scour depth (m) velocity (m/s)
d50 median grain size (mm) Wj dimensionless transport rate of the jth sediment
fj percentage (by volume) of the jth sediment fraction
fraction (x1 , x2 , x3 ) Cartesian coordinates (m)
G bedload transport function (˜x1 , ˜x2 ) mesh sizes in (x1 , x2 ) directions (m)
g acceleration due to gravity (¼ 9.81 m/s2 ) y normal distance from the channel bed to a
go reduced hiding factor velocity point (m)
(i, j, k) cell indices in (x1 , x2 , x3 ) directions yþ wall distance ( yu =)
J total number of sediment fractions z bed level above an arbitrary reference datum (m)
j jth sediment fraction or cell index in the x2 (˜z1 , ˜z2 ) difference in bed level between adjacent nodes in
direction the (x1 , x2 ) directions (m)
KM horizontal eddy viscosity (m2 /s) Æ angle between near-bed velocity and maximum
ks channel bed roughness height (mm) slope direction (degrees)
kþs roughness Reynolds number ( k s u =)  angle of steepest descent or maximum bed slope
m non-hydrostatic pressure correction (m2 /s2 ) (degrees)
n number of interior cells in the model channel  multiplication factor of c
(qb1 , qb2 ) bedload flux in (x1 , x2 ) directions (m3 /s)m o lower limit for 
s specific weight of sediments (Ł1 , Ł2 ) angle of slope of local bed surface in (x1 , x2 )
t time (s) directions (degrees)
te time to equilibrium (s) k von Karman constant (¼ 0.41)
˜t time step (s) º sediment porosity
(u1 , u2 , u3 ) Reynolds-averaged velocity components in  molecular kinematic viscosity of water (m2 /s)
(x1 , x2 , x3 ) directions (m/s) r density of water (¼ 1000 kg/m3 )

1
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

 coordinates in the vertical direction used the probabilistic approach to calibrate ds formulae and Hong
g geometric standard deviation of grain size et al. (2012) applied the support vector regression technique to
distribution predict time-dependent ds :
 arithmetic standard deviation of grain size
distribution The analyses mentioned above did not consider fractional trans-
b bed shear stress (N/m2 ) port of non-uniform sediments. At best, they indirectly considered
(b1 , b2 ) bed shear stress in (x1 , x2 ) directions (N/m2 ) non-uniformity through the mean grain size d50 along with the
c critical shear stress (N/m2 ) geometric standard deviation of grain size distribution g :
 argument of the function G Recently, Brandimarte et al. (2012) pointed out the importance of
c angle of repose of sediments (¼ 348) considering various aspects of the scour problem and improving
o normalised bed shear stress understanding of scour-triggering dynamics.
ø straining factor
Current understanding of the scour problem is based mostly on
1. Introduction laboratory observations. Experiments with uniform sediments
Bridge pier scour has been a major cause of many bridge failures. have shown a temporally varying scour depth ds and time to
For safe and cost-effective design of pier foundations, one needs equilibrium te (Melville and Chiew, 1999) and the effects of pier
accurate scour predictions for given hydraulic and geometric diameter D (Ettema et al., 2006) and D/d50 (Sheppard and Miller,
conditions. So far, this need has not been met. Typically, 2006) on ds : Some of the findings from these studies are
sediments at a natural riverbed are non-uniform. It is difficult to inconclusive: Sheppard and Miller (2006) suggested D/d50 having
accurately predict non-uniform sediment scour around a pier a progressively weaker influence on ds at stronger approach flow
(Hong et al., 2012) due to the complex three-way interaction relative to the threshold velocity for the initiation of sediment
involving the approach flow, pier obstruction and erodible channel motion, whereas Ettema et al. (2006) reported decreasing ds with
bed (Figure 1). The purpose of this work was to develop a three- increasing D.
dimensional morphodynamic model for scour simulations with
non-uniform sediments and to verify the model results using Based on experiments with non-uniform sediments (fine to
experimental data. medium size gravel), Raikar and Dey (2005) made the sugges-
tions that scour depth ds increases with decreasing gravel size,
Based on dimensional analysis, some investigators (Melville and and dimensionless time to equilibrium increases with increasing
Sutherland, 1988; Raudkivi, 1986) have dealt with the simplified pier Froude number for uniform gravel and decreases with
case of scour with uniform sediments. The outcome is semi- increasing g for non-uniform gravel. This implies that fractional
empirical formulae for determining the maximum value of scour transport should be considered in non-uniform sediment scour
depth (ds ) from input parameters including the approach flow predictions. With non-uniform sediments, ds is smaller than or
depth, velocity, pier diameter (D) and median sediment grain size equal to that with uniform sediments of the same d50 (Chiew and
(d50 ). The output is typically the normalised scour depth ds /D. Melville, 1989).
According to Lee and Strum (2009), the formulae generally
overpredict at the field scale. A few investigators have considered There are uncertainties in translating laboratory-based scour
non-uniform sediments in their scour analyses: Johnson (1992) predictions from the small laboratory scale to the field scale. In
this regard, numerical modelling has advantages – it can easily
be implemented at the field scale and is less time consuming and
Average approach flow velocity ⫽ 0·28 m/s less expensive than laboratory experiments.
Approach flow depth ⫽ 0·2 m
Pier diameter ⫽ 0·1 m
Downstream A review of the literature shows some modelling studies of pier
l
wal 5m
Side 1m
scour with uniform sediments (Baranya et al., 2013; Khosrone-
jad et al., 2012; Nagata et al., 2005; Olsen and Kjellesvig,
1998; Olsen and Melaaen, 1993; Roulund et al., 2005). The
5m
numerical model used by Olsen and Melaaen (1993) is based
on the steady-state Reynolds-averaged Navier–Stokes equations.
x3 x1 Such models save computational time but face uncertainties in
x2 predicting scour development, partly because unsteady terms
h eam are omitted from the equations. In the other studies, the
roac pstr
App from u developed models use the unsteady Reynolds-averaged Navier–
flow
Stokes equations along with complex turbulence closure
Figure 1. Three-dimensional view of the model channel, with a schemes. As a result, these models are of low computational
circular pier in the middle efficiency and thus have a limited potential for applications to
pier scour at the field scale. Many numerical studies focusing

2
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

on pier hydrodynamics in the absence of sediment transport ro K M (@u1 =@x3 , @u2 =@x3 ) ¼ (w1 , w2 ) ¼ 0
have been reported in the literature (e.g. Kirkil et al., 2008); a
review of recent progress on the topic of pier hydrodynamics is
given by Paik et al. (2010). Fluid particles on the surface will remain there all the time,
expressed as
The above-mentioned studies of pier scour with uniform sedi-
ments have ignored the effects of armouring, selective transport
and relative exposure on scour development. Chiew (1991) u3 ¼ u1 @=@x1 þ u2 @=@x2 þ @=@t
reported experimental evidence that, at low flow velocity, an
armoured bed of coarse particles can form with a slight transport
of fine particles, whereas at high velocity, particles of all sizes At the channel bed, the bed shear stress is calculated according to
participate in transport and a transition flatbed forms at all the quadratic law
sediment sizes. Numerical modelling of the indicated scour
process is difficult. Three-dimensional numerical models invol-
ro K M (@u1 =@x3 , @u2 =@x3 ) ¼ (b1 , b2 )
ving non-uniform sediments have been applied to channel bends
and lateral contractions (Bui and Rutschmann, 2010; Tritthart and
Schober, 2011). To the best of the authors’ knowledge, no The normal component of the velocity vector at the channel bed
application to pier scour with non-uniform sediments has been is zero, given by
reported in the literature.

Modelling methods that allow for fractional bedload transport u3 ¼ u1 @H=@x1  u2 @H=@x2
and particle hiding effects are described in Section 2. Section 3
presents the model results of non-uniform sediment scour around
a circular pier and a comparison with experimental data. Conclu- where H is the depth of the flow from datum. The channel and
sions are drawn in Section 4. pier sidewalls are fully slippery; the velocity component normal
to the sidewalls is zero. At the upstream open boundary, steady
approach flow is prescribed; at the downstream open boundary,
2. Methodologies
water surface elevation is given.
2.1 Hydrodynamics equations and boundary
conditions 2.2 Hydrodynamics solution techniques
Computations of the velocity field and free surface elevation use The model equations are transformed to the -coordinate in the
the non-hydrostatic hydrodynamics model of Pournazeri et al. vertical. The free water surface and the channel bed are trans-
(2013), which is an extension of an existing hydrostatic model formed to coordinate planes. This terrain-following feature of the
(HydroQual, 2002). The extension involves incorporating a new -coordinates has advantages in simulating scour development.
computational module of non-hydrostatic pressure corrections Multi-layer modelling has advantages of reducing computational
(Casulli, 1999) into the hydrostatic model, which is necessary for costs. The transformed model equations are then cast into finite
pier scour applications because significant vertical acceleration is difference equations and are numerically solved. The finite
expected in a scour hole. In the hydrodynamics model, the difference techniques used have second-order accuracy in space
governing equations are the Reynolds-averaged equations of and time.
continuity and momentum balance. Turbulence closure is ob-
tained using the scheme of Smagorinsky (1963) in the horizontal For better computational efficiency, mode splitting techniques are
to parameterise the effect of sub-grid motions on the flow field used to separate vertically integrated flow (external mode) and
resolved by the horizontal grids and the second-order scheme of flow structure in the vertical (internal mode). This technique
Mellor and Yamada (1982) in the vertical to allow for the effect allows computation of the free water surface with little sacrifice
of vertical shear on momentum mixing. Details about theoretical in time. The staggered computation grid is used due to its high
considerations, solution procedures, empirical relationships and effectiveness. Solution procedures start with solving the external
calibration constants related to the two turbulence closure and internal modes without non-hydrostatic pressure for provi-
schemes are also provided (Mellor and Yamada, 1982; Smagor- sional velocity components, and then make non-hydrostatic
insky, 1963). In the vertical, the water column is divided into a corrections to the velocity field.
number of layers and the governing equations are integrated over
the thickness of each layer in order to obtain model equations for 2.3 Non-hydrostatic pressure corrections
layer-averaged velocity components. The boundary conditions are Suppose that the model domain is discretised into cells with
as follows. dimensions (imx, jmx, kmx) in the (x1 , x2 , x3 ) directions (Figure
1). It can be shown that the non-hydrostatic pressure correction m
At the free surface, the wind shear stress w is zero, given by is governed by a seven-diagonal linear system of the form

3
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

2 32 3 2 3
A1 B1 0 : C1 0 : G1 0 : : 0 m1 R1
6 76 7 6 7
6 D2 A2 B2 0 : C2 0 : G2 : : : 76 m2 7 6 R2 7
6 76 7 6 7
6 0 D3 A3 B3 : : C3 : : : : : 7 6 7 6 7
6 76 : 7 6 : 7
6 76 7 6 7
6 : 0 : : : : : : : : : 0 7 6 7 6 7
6 76 : 7 6 : 7
6 76 7 6 7
6 Eimx1 : 0 : : : : : : : : Gnm2 7 6 7 6 7
6 76 : 7 6 : 7
6 76 7 6 7
6 0 Eimx : 0 : : : : : : : 0 7 6 7 6 7
6 76 : 7 ¼ 6 : 7
6 76 7 6 7
6 : 0 : : : : : : : : : : 7 6 7 6 7
6 76 : 7 6 : 7
6 76 7 6 7
6 F np1 : 0 : : : : : : : : C nm1 76 : 7 6 : 7
6 76 7 6 7
6 76 7 6 7
6 : F np2 : : : : : : : : : : 76 : 7 6 : 7
6 76 7 6 7
6 : : : : : : : : : : : 0 7 6 7 6 7
6 76 : 7 6 : 7
6 76 7 6 7
6 : : : : : : : : : : : Bn1 7 6 7 6 7
4 54 : 5 4 : 5
1: 0 0 : 0 Fn 0 : En : : 0 Dn An mn Rn

with np1 ¼ (imx  2)( jmx  2) þ 1, np2 ¼ (imx  2)( jmx  2) using up in the (x1 , x2 ) directions gives the bed shear stress
þ 2, nm2 ¼ n  (imx  2)( jmx  2) and nm1 ¼ n  (imx  2): (b1 , b2 ) in the corresponding directions. The law of the wall is
The unknowns or non-hydrostatic pressure values for the interior valid for the wall distance yþ . 11:53: The actual y þ values are
cells are m1 , m2 , . . ., m n : A1 , B1 , . . ., D n , A n are the coefficients determined from the shortest distance from the channel bed
of the system and R1 , R2 , . . ., Rn are the constant terms. surface to the location of up in question. This study uses
Subscripts 1 and n correspond to the first interior cell velocities from near-bed locations with 300 , yþ , 500, as
(i, j, k) ¼ (2, 2, 2) and the last interior cell (i, j, k) ¼ (imx  1, opposed to the bottom-layer velocity. The b results are more
jmx  1, kmx) respectively. The system is symmetrical and has accurate and contain less noise (Pournazeri et al., 2013).
positive definition. It is solved for the unknowns through
preconditioned conjugate gradient iterations. Suppose that a sediment mixture is divided into J size fractions,
and the jth fraction has a percentage of f j by volume. The
Boundary conditions on m are as follows. On the sidewalls and percentages of individual fractions are updated at each time step
pier surface, m is such that the normal flux is zero. At open during a simulation. For the jth fraction, the per-unit width
boundaries, m is specified. At the water surface, m is zero. Once bedload rates in the (x1 , x2 ) directions are given by
m is solved, the velocity components (u1 , u2 , u3 ) and the free
: :
surface elevation for all interior cells can be obtained (details are (1b15 , 1b25 )
given by Pournazeri et al. (2013)). The velocity components are (qb1 , qb2 ) ¼ f Wj
2: r1:5 (s  1)g j
used in calculations of bedload.

2.4 Fractional bedload transport The transport rate W j is related to a transport function G through
The bed shear stress b is calculated using the logarithmic law of W j ¼ 0:00218G(): The functional form of G() is given by
the wall Parker (1990). Its argument, , combines the effects of the
normalised bed shear stress o , the particle reduced hiding factor
:
up =u ¼ k1 ln(yþ =k þ :
s )þ85
go and a straining factor ø as  ¼  g  ø, g ¼ (d j =d g )0 0951
and

with b related to the friction velocity u as b ¼ ru2 : Note that (b1 , b2 )
 ¼
use of the logarithmic law (as opposed to more complicated non- 3: (s  1)rgd g c
equilibrium wall laws) is sufficient since no severe pressure
gradients are expected to occur. The roughness height ks is taken
as 3d50 : This choice is within the range of values suggested by a The straining factor ø allows for the effect of the difference
large number of studies, including those of Camenen et al. between substrate and surface grain size distributions. Data fitting
(2006), Wu et al. (2000) and van Rijn (1984). The suggestion is (Parker, 1990) yields the relationship
that ks is of the order of d50 or of some larger grain size
percentiles for a mobile bed; commonly, ks is taken as equal to ø ¼ 1 þ   (ø  1)= 
one to five times d50 , d65 , d84 or d90 : Applying the law of the wall

4
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

in which considered conceptually as the active layer. Note that the numer-
ical techniques described in this paper do not require a prescribed
:
ø ¼ 0:45 þ 1:02e0 6 active-layer thickness as an additional model parameter.
:
  ¼ 1:5  1:67e0 9 During scour development, the slope of the channel bed surface
X 0:5 needs to be examined to ensure that the slope resulting from bed-
 ¼ ln2 (d j =d g ) ln2 2 f j level change (Equation 5) does not exceed the angle of repose of
j sediments. The slope angle in the (x1 , x2 ) directions is calculated
as tan(Ł1 , Ł2 ) ¼ (˜z1 =˜x1 , ˜z2 =˜x2 ): At any time, if bed-level
change results in Ł1 . c or Ł2 . c , an adjustment at two
The critical shear stress c in Equation 3 is specified in two steps. neighbouring nodes will be made by lowering the higher bed
First, an appropriate value is determined based on the mean grain level and raising the lower one, given by
diameter. Then, the value is multiplied by a modification factor 
to allow for the effect of bed slope (Roulund et al., 2005) since the ˜zj ¼ 0:5(tan Łj  tan c )˜xj j ¼ 1, 2
shear stress required to initiate sediment movement on a sloping
bed decreases with increasing slope. This factor is given by
To prevent small-scale noise in the bed level, which can possibly
:
 ¼ max[cos (1  sin2 Æ tan2 = tan2 c )0 5 cause numerical instability, the bed level is smoothed using a
five-point median filter (Liang and Cheng, 2005).
4:  cos Æ sin = tan c ,  ]
2.6 Model run setup
Model runs were set up for the same hydraulic conditions and
A lower limit, o , has been imposed on  because the first pier geometry as an experiment (run S3) reported by Chang et al.
expression in the parentheses approaches zero at a slope angle (2004). The model channel is 10 m long and 1 m wide, with a
equal to the angle of repose of sediments, and this would cause circular pier of 0.10 m diameter located in the middle (Figure 1).
infinite bedload and hence numerical instability. The lower limit At the upstream boundary, the approach flow is steady, with a
is thus used to avoid instability. depth of 0.2 m and a prescribed vertical profile of u1 velocity,
which gives a discharge of 0.056 m3 /s. The channel has a
2.5 Morphologic update longitudinal slope of 0.0007.
Bed-level changes caused by bedload transport (Equation 2) are
calculated using the Exner equation The experiment of Chang et al. (2004) used bed sediments with a
median grain size d50 of 1 mm and a geometric standard
@z XJ deviation  g ¼ 2: From these two parameters, a grain size
(1  º) ¼ (@qb1 =@x1 þ @qb2 =@x2 )j distribution of six fractions was constructed. The grain sizes used
5: @t j¼1
by Chang et al. (2004) were 0.19, 0.38, 0.75, 1.5, 3 and 6 mm
(fine sands to fine gravel) and the corresponding percentages were
7, 22, 45.5, 17.5, 5 and 3% (Figure 2). For the indicated sediment
Equation 5 is cast into finite difference form using forward
stepping in time and upwind differencing in space. A live layer of 100
thickness of 2d50 is assumed; this thickness value is used merely
as the upper limit on bed-level changes computed at any one time
80
step during a pier scour simulation. Over the simulation time
period, the bed level is updated continuously until a state of
Percent finer: %

equilibrium is reached. Note that setting a limit on bed-level 60


changes at any one time step is for numerical stability considera- d50 ⫽ 1 mm
tions. It has some influence on computational time to complete a σg ⫽ 2
40
simulation and reach a state of equilibrium, but does not affect
equilibrium solutions.
20
Some researchers (Armanini and Di Silvio, 1988) related active-
layer thickness to flow depth, dune height or grain size. There is 0
0 2 4 6
no consensus on how active-layer thickness is related to physical
d: mm
processes taking place in the layer and how to define the thickness
(Bui and Rutschmann, 2010). Following Bui and Rutschmann Figure 2. Sediment grain size distribution
(2010), the maximum bed deformation at the current time step is

5
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

grain sizes, the critical shear stress in Equation 3 is estimated to 2.7 Model accuracy
be 0.25 Pa , c , 0.75 Pa (Paphitis, 2001) and the corresponding The accuracy of hydrodynamics predictions using the techniques
Shields parameter ranges from 0.015 to 0.046 (note that c is a described was validated by Pournazeri et al. (2013) by comparing
calibration parameter). velocity predictions with available laboratory measurements (Graf
and Istiarto, 2002). The predictions were obtained under the same
In the horizontal, the channel is covered with non-uniform mesh conditions of channel geometry, pier configuration and approach
(Figure 1). The mesh is as fine as 0.01 m for the near-pier area to flow as in the laboratory experiments. As illustrated in Figure 3
resolve velocity gradients. Away from the pier, the mesh increases of Pournazeri et al. (2013), the predicted and measured velocities
in size gradually by a factor less than or equal to 1.4 between for a series of locations at the vertical cross-section along the
neighbouring cells, which helps to save computational time. The channel centreline are in reasonable agreement, especially with
coarsest mesh is 0.098 m in size. The aspect ratio of mesh sizes regard to strong velocities. Accurate predictions of velocity for
in the (x1 , x2 ) directions is limited to 9.8 or lower to avoid flow locations of strong flow are the most important for calculations of
distortion. the bed shear stress and hence bedload transport and scour depth.
More details about the hydrodynamics validation are given by
In the vertical, the water column is divided into seven layers of Pournazeri et al. (2013).
varying thickness. The layer thickness is small near the bottom
and relatively large near the water surface. The thickness is 3% The hydrodynamics performance demonstrated by Pournazeri et
of the local flow depth for the bottom layer and 23% for the al. (2013) is considered to be relevant to this study for two simple
surface layer, which correspond to 0.006 m and 0.046 m respec- reasons. Firstly, this paper is an extension of Pournazeri et al.
tively, at the beginning of a simulation. The use of a small layer (2013) in the morphodynamics module from uniform to non-
thickness for the lower water column helps resolve the bottom uniform sediments; the hydrodynamics module is the same.
boundary layer, whereas use of a relatively large layer thickness Secondly, the application covered in this paper closely resembles
for the upper water column increases computation efficiency, with that covered by Pournazeri et al. (2013) in terms of channel
minimal effects on scour calculations. geometry, pier configuration and approach flow.

Model runs start from the initial condition of a flat channel bed
3. Results and discussion
and continue until a state of equilibrium. The time step used in
the model (˜t ¼ 0.002 s) satisfies the stability constraints for the 3.1 Sensitivity test
external and internal modes. Key parameters and their values are Three runs (P1, P2 and P3) were carried out to test the model
listed in Table 1. with regard to numerical stability, solution convergence and

Approach flow depth: m 0.20


Cross-sectionally averaged approach flow velocity uo : m/s 0.28
Approach flow Froude number 0.2
Geometric mean grain size dg : mm 0.7
Sediment porosity º 0.47
Ratio of approach flow depth to pier diameter 2
Ratio of pier diameter to d50 100
Mesh dimension in x1 direction (150, 155, 170)
Mesh dimension in x2 direction (36, 43, 46)
Mesh size in x1 direction ˜x1 : cm (1–9.8, 0.9–9.6, 0.8–8.3)
Mesh size in x2 direction ˜x2 : cm (1–5, 0.9–3.5, 0.8–3.2)
Number of layers in x3 direction 7
Maximum ˜x1 /˜x2 (9.8, 10.6, 10.37)
Simulation period: s (2290, 550, 550)
Bottom drag coefficient 0.0025
Constant in calculation of horizontal mixing coefficient 0.1
Criterion for convergence 106
Ratio of uo to armour velocity 0.45

Table 1. Summary of model variables and parameters; the three


different values or ranges in parentheses for some of the
parameters correspond to runs P1, P2 and P3 respectively

6
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

Run P1
independence of the results on mesh configuration. These three 1·5
runs differed in mesh sizes, ˜x1 and ˜x2 (Table 1). In the x1 ds /D
direction, ˜x1 ranged from 1.0 cm to 9.8 cm (or 0.1D to 0.98D) 1·0
for run P1, and was progressively refined to 0.9–9.6 cm for run 0·2
P2 and 0.8–8.3 for run P3. In the x2 direction, the mesh size was 0·5
also refined from run P1 to P2 and to P3. The aspect ratio ˜x1 / 0·1

x2 /D
˜x2 was similar for the three runs. Details about the setup are 0
given in Table 1. 0
⫺0·5
⫺0·1
Bed-level changes around the pier at a selected model time of
⫺1·0
550 s for runs P1, P2 and P3 are compared in Figure 3. The
⫺0·2
changes for P1 and P2 are similar. For P3, scour depth and
⫺1·5
deposition height are slightly larger than the predictions for P1 ⫺1·5 ⫺1·0 ⫺0·5 0 0·5 1·0 1·5
and P2. All three runs predicted two sites of scour upstream of
Run P2
the pier’s centre and two sites of deposition downstream. The 1·5
scour is deeper on the site to the right of the pier (to an ds /D
observer facing downstream) than to the left; this is true for all 1·0
the runs. The maximum values of scour depth for runs P1, P2 0·2
and P3 were 0.254D, 0.256D and 0.290D respectively and the 0·5
maximum deposition heights for the three runs were 0.278D, 0·1
x2 /D

0.264D and 0.253D (Table 2). Table 2 shows comparable 0


results of scour/deposition areas as well as their central 0
positions for the three runs, meaning that the model results are ⫺0·5
⫺0·1
independent of mesh configuration. Runs P1, P2 and P3 used
⫺1·0
seven layers in the vertical. To further test if the model results
⫺0·2
are sensitive to layer resolution, the number of layers was
⫺1·5
increased from 7 to 14 and P3 was re-run. The differences in ⫺1·5 ⫺1·0 ⫺0·5 0 0·5 1·0 1·5
near-bed flow and bed-level change (not shown) were insignif-
Run P3
icant between the 7-layer and 14-layer runs. The mesh used in 1·5
run P1 was the most efficient and was thus used for subsequent ds /D
runs. 1·0
0·2
3.2 Equilibrium solution 0·5
0·1
A series of calibration runs was carried out using the critical
x2 /D

shear stress c in Equation 3 and the limit o in Equation 4 as 0


0
calibration parameters. The values tested were in the range
0:25 < c < 0:32 Pa and 0:1 <  < 0:15: A value of 0.25 Pa for ⫺0·5 ⫺0·1
c is suitable (Paphitis, 2001) for a sediment mixture with
⫺1·0 ⫺0·2
d50 ¼ 1 mm; the corresponding Shields parameter is 0.015. The
results from two of the calibration runs (C1 and C2) are
⫺1·5
described. For run C1, with  ¼ 0:1, and run C2, with ⫺1·5 ⫺1·0 ⫺0·5 0 0·5 1·0 1·5
 ¼ 0:15, the predicted values of scour depth ds at the pier nose x1 /D
(or the maximum ds at the location next to the leading edge of
the pier) are compared with the measurements of Chang et al. Figure 3. Contours of bed-level change at a selected model time
(2004) as a time series in Figure 4. During the initial state of of 550 s, showing scour (ds , 0) and deposition (ds . 0) for runs
scour (t=te , 0:2), the bed-level changes rapidly, giving rise to a PI, P2 and P3
scour hole around the pier. Subsequently, the scour hole grows
gradually with time (t=te . 0:2).
reaches a state of equilibrium at a model time of 38.16 min and
As shown in Figure 4, at time to equilibrium te , the measurements ds at the pier nose is 0.458D, with a prediction error as low as
of Chang et al. (2004) give a scour depth of 0:43D: For run C1, 6.5%. The scour depth reaches 0.471D as the maximum; this
scour reaches a state of equilibrium at a model time of 35 min occurs at a location very close to the nose (x1 =D ¼ 0:7,
and the maximum ds is 0.535D, which occurs at the location x2 =D ¼ 0:25). For run C2, ds predictions at the maximum scour
(x1 =D ¼ 0:57, x2 =D ¼ 0:86). At the pier nose, ds ¼ 0.48D; location and at the pier nose agree well with the laboratory
the prediction error is slightly lower than 12%. For run C2, scour measurements.

7
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

Run Maximum Scour depth . 0.10D Maximum Deposition height . 0.10D


scour deposition
depth Area Central position (x1 /D, x2 /D) height Area Central position (x1 /D, x2 /D)

P1 0.254 0.39 (0.57, 0.85) 0.278 0.36 (0.99, 0.87)


P2 0.256 0.41 (0.67, 0.89) 0.264 0.33 (0.97, 0.83)
P3 0.290 0.40 (0.55, 0.90) 0.253 0.34 (0.95, 0.85)

Table 2. Between-run comparison of predicted scour and


deposition. The scour depths and deposition heights were
normalised by D and the areas by D2 /4

0·01 Next to upstream nose


Change of ds /D

(Run C2)

1·0 ⫺0·01
0·8 1·0
t /te

e
r sit t)
0·8 x . scou edimen
Ma orm s
Measurement f
(uni
Normalised scour depth ⫺ds /D

(Chang et al., 2004)

0·6

0·4

Max. scour site (run C1)


0·2
Max. scour site (run C2)

Next to upstream nose (run C1)


0 Next to upstream nose (run C2)

0 0·2 0·4 0·6 0·8 1·0


Normalised time t /te

Figure 4. Time series of scour depth ds at the maximum scour site after normalised time t/te ¼ 0.8. The difference fluctuates with a
and/or around the pier nose for runs C1 and C2, and for a very small amplitude (,0.6%) and diminishes towards zero (zero
calibration run with uniform sediments. The inset shows the difference means that the scour depth reaches a state of
difference in normalised scour depth (around the pier nose for run equilibrium)
C2) between the present time and preceding model output time

For comparison, the local or maximum ds for the hydraulic and overprediction compared to the measurement of Chang et al.
pier geometric conditions given in Table 1 was calculated using a (2004) (0.43D). The calculations used a Froude number of 0.2 and
widely used empirical formula. Johnson’s equation for maximum a sediment gradation of 3.1, as given in Chang et al. (2004). In
ds (Johnson, 1992) along with coefficients from Muzzammil and this modelling study, the predictions for run C2 had an accuracy
Siddiqui (2009) gives a value of 0.878D. This is a significant better than 90% and results from this run are discussed further.

8
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

0·35 m/s
A time series of ds for a calibration run with uniform sediments
(d50 ¼ 1 mm) is shown as the thin solid curve in Figure 4. This 4·5
calibration run used the same conditions as C2 with non-uniform
sediments (d50 ¼ 1 mm, g ¼ 2), but the sediments were of 3·0
uniform size. This calibration run predicted a maximum ds at
equilibrium almost double that of run C2. The implication is that 1·5
sediment non-uniformity plays an important role in scour

x2/D
development. Specifically, non-uniformity leads to a reduction in 0
the magnitude of pier scour. This is consistent with the finding of
Chiew and Melville (1989) who reported a reduction in scour ⫺1·5
hole depth depending on the value of g : This reduction can
possibly be explained by the effect of armouring developed on ⫺3·0
non-uniform bed sediments.
⫺4·5
3.3 Velocity field in the horizontal ⫺4·5 ⫺3·0 ⫺1·5 0 1·5 3·0 4·5 6·0
The velocity vector field at a state of equilibrium for run C2 is (a)
shown in Figure 5. The approach flow has a cross-sectionally 0·35 m/s
averaged velocity uo of 0.28 m/s (Table 1). In the surface layer
4·5
(Figure 5(a)), as the flow approaches the pier, the longitudinal
velocity decreases from u1 ¼ 0.37 m/s at the upstream edge of
the scour hole to u1 ¼ 0.15 m/s close to the pier. The transversal 3·0
velocity u2 is relatively high on the upstream side of the pier,
with a maximum value of 0.22 m/s. The longitudinal velocity is 1·5
as large as u1 ¼ 0.41 m/s on both sides of the pier. In the wake
x2/D

region, the flow direction reverses and results in weak circulation 0


clockwise and counter-clockwise. Far away from the pier towards
downstream (x1 =D . 16:2), u1 recovers the maximum value of ⫺1·5
0.35 m/s, whereas u2 decreases. Accordingly, the wake intensity
weakens. ⫺3·0

In the second layer from the bottom (Figure 5(b)), as the flow ⫺4·5
⫺4·5 ⫺3·0 ⫺1·5 0 1·5 3·0 4·5 6·0
approaches the pier, u1 decreases from 0.2 m/s to almost zero at
x1/D
the upstream edge of the scour hole. Then, flow reversal takes
(b)
place, with a flow magnitude of 0.13 m/s in the vicinity of the
pier. On the upstream side of the scour hole, u2 is strong, with a Figure 5. Velocity field for run C2 at the surface layer (a) and at
magnitude of 0.27 m/s. The same holds true for u1 , with a the second layer above the bed (b); every second column and row
magnitude of 0.24 m/s. The resultant flow velocity has sufficient are plotted
strength to carry sediments away from the upstream side towards
downstream. Downstream of the pier, the flow reverses its
direction, with a magnitude of 0.08 m/s as the maximum, and
produces wakes circulating clockwise and counter-clockwise. s), occurring at a depth equal to 55% of the total depth below the
Further away from the pier towards downstream, u1 increases to water surface. Immediately downstream of the pier, the flow is
almost 0.14 m/s, with reduced wake intensity. downward; the flow remains downward over a horizontal distance
of x1 =D ¼ 1:5 from the pier. Further downstream from the pier,
3.4 Flow–sediment interaction there is an increase in horizontal flow velocity. The vertical flow
During scour development, the velocity field, pier and channel velocity decreases to almost zero.
bed interact. To illustrate this interaction, vertical cross-sections
are plotted in Figure 6, showing velocity vectors and channel bed Towards the end of the initial stage of the scour process (Figure
profiles at selected time steps. The cross-sections are located near 6(b)), eddy motions upstream of the pier intensify, especially at the
the channel centreline (at x2 =D ¼ 0:15). Shortly after the begin- foot of the pier. As the flow approaches the pier, u1 decreases by
ning of run C2 (Figure 6(a)), as the flow approaches the pier, the 50% (from 0.37 m/s in the approach channel) and the flow reverses
longitudinal velocity u1 decelerates by 72% (from 0.32 m/s in the direction near the channel bed. The reversal flow has a maximum
approach channel), which creates a significant downflow just speed of 0.16 m/s. At the pier nose, the downflow has a maximum
upstream of the pier. The downflow has a maximum velocity of speed of 0.83 m/s, at a depth equal to 72.5% of the total depth below
0.22 m/s (close to the average approach flow velocity of 0.28 m/ the water surface. Immediately downstream of the pier, the flow is

9
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

1·5 1·5

0 0
x3 /D

⫺1·5 ⫺1·5

⫺3·0 0·35 m/s ⫺3·0 0·35 m/s

⫺4·5 ⫺4·5
⫺5 ⫺4 ⫺3 ⫺2 ⫺1 0 1 2 3 4 5 6 ⫺5 ⫺4 ⫺3 ⫺2 ⫺1 0 1 2 3 4 5 6
(a) te /t ⫽ 0·004 (b) te/t ⫽ 0·262

1·5 1·5

0 0
x3/D

⫺1·5 ⫺1·5

⫺3·0 0·35 m/s ⫺3·0 0·35 m/s

⫺4·5 ⫺4·5
⫺5 ⫺4 ⫺3 ⫺2 ⫺1 0 1 2 3 4 5 6 ⫺5 ⫺4 ⫺3 ⫺2 ⫺1 0 1 2 3 4 5 6
x1/D x1/D
(c) te /t ⫽ 0·524 (d) te /t ⫽ 1

Figure 6. Development of the flow field and bed scour at vertical


cross-sections (at x2 =D ¼ 0:15) for run C2

downward, with a maximum value of 0.1 m/s. The flow becomes sediments from the foot of the pier. Downstream of the pier, eddy
upward at a horizontal distance of x1 =D ¼ 2:4: Further downstream motions are essentially the same as in Figure 6(c).
of the pier, the flow becomes parallel to the channel bed and u1
increases to a maximum of 0.39 m/s. Eddy motions downstream of Upstream of the pier, the bed slope along the channel centreline
the pier are weak and are not sufficient to cause scouring. reaches 318 (Figure 6(d)), which is close to the angle of repose of
uniform sediments (348). In principle, the angle of repose of non-
During the gradual development stage of scour (Figure 6(c)), the uniform sediments can be calculated using the formulation of
approach flow velocity drops by 50% in the surface layer. The Yang et al. (2009). However, this requires tracking of individual
flow reverses its direction near the channel bed, with a maximum sediment grains, which is not feasible. As an approximation, the
speed of 0.16 m/s. At the pier nose, the downflow has a maxi- angle of repose of non-uniform sediments is taken as 348, which
mum speed of 0.65 m/s, occurring at a depth equal to 72.5% of corresponds to uniform sands with d50 ¼ 1 mm.
the total depth below the water surface. Immediately downstream
of the pier, downflow of up to 0.12 m/s takes place. The 3.5 Scour around the pier
longitudinal velocity is the reverse direction over a horizontal The development of scour (ds , 0) and deposition (ds . 0) is
distance of x1 =D ¼ 1:5: Far downstream of the pier, the flow is shown in Figure 7. At the initial stage of scour development
parallel to the bed with a maximum value of 0.37 m/s. (Figure 7(a)), scouring emerges from two side areas upstream of
the pier that are quite symmetric about the channel centreline
At a state of equilibrium (Figure 6(d)), the longitudinal velocity through the pier centre. The maximum value of scour depth is
decreases by 50% in the surface layer. Near the channel bed, the 0.168D, located at (x1 =D ¼ 0:7, x2 =D ¼ 0:35). At this stage,
flow reverses its direction, with a maximum speed of 0.13 m/s. At the pier nose has not shown any significant scour. Downstream of
the pier nose, the downflow has a maximum speed of 0.75 m/s at the pier, sediment deposition begins on both sides of the channel
a depth equal to 72.5% of the total depth below the free surface. centreline, with a maximum height of 0.27D located at
This maximum value is 2.39 times the average approach flow (x1 =D ¼ 0:97, x2 =D ¼ 0:86).
velocity. This value is an overprediction. Copp and Johnson
(1987) have suggested 0.8 times. The overprediction is probably Near the end of the initial stage (Figure 7(b)), ds increases to
due to the use of the fully slippery condition at the pier surface. 0.269D near the pier nose at (x1 =D ¼ 0:69, x2 =D ¼ 0:25).
Upstream of the pier, eddy motions are strong, removing Deposition has a maximum height of 0.272D located at

10
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

ds /D ds /D

0·2
0·2 1·0
1·0
0·1
0·5 0·5
0·1
0
x2/D

0 0
0
⫺0·5 ⫺0·5 ⫺0·1

⫺1·0 ⫺0·1 ⫺1·0 ⫺0·2

⫺1·0 ⫺0·5 0 0·5 1·0 1·5 2·0 ⫺1·0 ⫺0·5 0 0·5 1·0 1·5 2·0
(a) t /te ⫽ 0·13 (b) t /te ⫽ 0·263

ds /D ds /D

0·2 0·2
1·0 1·0

0·5 0·5
0 0
x2/D

0 0

⫺0·5 ⫺0·5 ⫺0·2


⫺0·2

⫺1·0 ⫺1·0
⫺0·4
⫺0·4
⫺1·0 ⫺0·5 0 0·5 1·0 1·5 2·0 ⫺1·0 ⫺0·5 0 0·5 1·0 1·5 2·0
x1/D x1/D
(c) t /te ⫽ 0·66 (d) t /te ⫽ 1

Figure 7. Contours of bed-level change for run C2

(x1 =D ¼ 2:67, x2 =D ¼ 0:86). The depth of the scour hole at the in grain size distribution versus run time from the initial values
pier nose increases to 0.205D. of fractional percentages (7, 22, 45.5, 17.5, 5 and 3% for the six
fractions respectively) were analysed. As an example, at a model
During the stage of gradual scour development (Figure 7(c)), ds time of 17 min or nearly halfway towards a state of equilibrium,
increases, the maximum value being 0.423D at approximately the the grain sizes changed to the following distribution.
same location as in Figure 7(b). At the pier nose, ds increases to
0.361D. Downstream of the pier, sediment deposition reaches a j The first fraction of sediment (grain size ¼ 0.19 mm) shows
maximum height of 0.372D, with the location shifting slightly maxima of about 80% near the nose of the pier and 10% at
towards downstream. The scour patterns in Figure 7(c) appear to the flanks of the pier. These percentages are substantially
be less symmetric about the channel centreline than those in different from the initial value of 7%.
Figure 7(b), possibly due to truncation and runoff errors. j The second fraction of sediment (grain size 0.38 mm) shows
maxima of 50% near the nose of the pier and 25% at the
At a state of equilibrium (Figure 7(d)), ds increases to a flanks.
maximum of 0.471D at the same location as in Figures 7(b) and j The third fraction of sediment (grain size 0.75 mm) shows a
7(c). At the upstream nose, ds reaches 0.458D. Downstream of maximum of 50% around the pier.
the pier, deposition shifts towards downstream, with a maximum j The fourth fraction of sediment (1.5 mm) shows maxima of
height of 0.4D near the end of the channel. Downstream of the 80% in a small area close to the pier and 20% at the flanks.
pier, especially along the centreline of the channel, the scour j The fifth fraction of sediment (3 mm) shows maxima of about
depth is very small due to weak vortices and wakes in the region. 80% close to the pier and 10% at the flanks.
The outer shape of the scour hole upstream of the pier is j The sixth fraction of sediment (6 mm) shows a small change
approximately a half cone, which is very similar to the results in percentage around the pier from its initial percentage.
reported by Richardson and Wacker (1999).
These changes to the grain size distribution indicate that the
Run C2 was redone with 14 layers in the vertical and the change bottom surface is coursing close to the nose of the pier inside the

11
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

scour hole. Sediment grains of relatively smaller sizes are more reality, it would be interesting to consider cohesive sediments,
mobile and are transported out of the scour hole towards down- suspended loads and unsteady flow conditions in future studies.
stream. Sediment grains of the smallest size account for a high
percentage in locations close to the pier, possibly due to the effect Acknowledgements
of particle hiding. Downstream of the pier, the grain size Financial support from NSERC through Discovery Grants held by
distribution has not changed significantly from the initial percen- S. S. Li and Concordia Research Chair grants held by F. Haghighat
tages. is acknowledged.

4. Conclusion
A three-dimensional morphodynamic model was developed for REFERENCES
predictions of pier scour with non-uniform sediments. The model Armanini A and Di Silvio G (1988) A one dimensional model for
has successfully been applied to predict the flow around a circular the transport of a sediment mixture in non-equilibrium
pier and the development of a scour hole. The predicted scour conditions. Journal of Hydraulic Research 26(3): 275–292.
depth at equilibrium agrees well with measurements reported by Baranya S, Olsen NRB, Stoesser T and Sturm TW (2013) A nested
Chang et al. (2004). grid based computational fluid dynamics model to predict
bridge pier scour. Proceedings of the Institution of Civil
For reliable predictions of pier scour with a sediment mixture, it Engineers –Water Management 167(5): 259–268, http://
is appropriate to consider selective transport and relative exposure dx.doi.org/10.1680/wama.12.00104.
of sediment particles. It is not sufficient to represent grain size Brandimarte L, Paron P and Baldassarre G (2012) Bridge pier
non-uniformity merely through an effective grain size; the scour: a review of processes, measurements and estimates.
consequences include a significant overprediction of equilibrium Environmental Engineering and Management Journal 11(5):
scour depth. 975–989.
Bui MD and Rutschmann P (2010) Numerical modelling of non-
Pier obstruction causes the approach flow to decelerate in the equilibrium graded sediment transport in a curved open
longitudinal direction and causes flow separation on the lateral channel. Computers & Geosciences 36(6): 792–800.
sides of the pier. Flow deceleration is compensated for by the Camenen B, Bayram A and Larson M (2006) Equivalent
creation of a strong downflow near the pier nose and vortices at roughness height for plane bed under steady flow. Journal of
the foot of the pier on the upstream side. The downflow, Hydraulic Engineering ASCE 132(11): 1146–1158.
combined with vortex motions, effectively removes bed sediments Casulli V (1999) A semi-implicit finite difference method for non-
from the upstream side. The downflow intensifies as scour hydrostatic, free-surface flows. International Journal for
develops. Its maximum intensity occurs at about 75% of the flow Numerical Methods in Fluids 30(4): 425–440.
depth below the free surface. Chang WY, Lai JS and Yen CL (2004) Evolution of scour depth at
circular bridge piers. Journal of Hydraulic Engineering ASCE
Scouring with non-uniform sediments emerges from the lateral 130(9): 905–913.
sides of the pier, caused by high velocities of flow bifurcating Chiew YM (1991) Bed features in non-uniform sediments.
around the pier nose. Scour patterns migrate towards the pier Journal of Hydraulic Engineering ASCE 117(1): 116–120.
nose as the local downflow and vortex motions intensify with Chiew YM and Melville BW (1989) Local scour at bridge piers
time. The initiating and migrating features are similar to those with non-uniform sediments. Proceedings of the Institution of
seen with uniform sediments (Pournazeri et al., 2013). At Civil Engineers – Water Management 87(2): 215–224.
equilibrium, the maximum slope angle of the bed surface on the Copp HD and Johnson JP (1987) Riverbed Scour at Bridge Piers.
upstream side of the scour hole is slightly smaller than the angle Department of Civil and Environmental Engineering,
of repose of sediments. Downstream of the pier, flow circulations Washington State University, Pullman, WA, USA, Report
in the vertical and horizontal produce relatively weak vortices WA-RD-118.1.
and wakes, whose intensities decrease with increasing distance Ettema R, Kirkil G and Muste M (2006) Similitude of large-scale
towards downstream. turbulence in experiments on local scour at cylinders. Journal
of Hydraulic Engineering ASCE 132(1): 33–40.
The model presented here is of practical relevance due to the use Graf W and Istiarto I (2002) Flow pattern in the scour hole
of the multi-layer modelling approach combined with non- around a cylinder. Journal of Hydraulic Research 40(1):
hydrostatic pressure corrections. The model is much more 13–20.
computationally efficient than other pier scour models that solve Hong J, Goyal MK, Chiew YM and Chua L (2012) Predicting
the primitive Navier–Stokes equations. Furthermore, it can easily time-dependent pier scour depth with support vector
be applied at the field scale, without scaling effect uncertainties, regression. Journal of Hydrology 468–469: 241–248.
and its capability to properly handle non-uniform sediments HydroQual (2002) A Primer for ECOMSED, Version 1.3.
ensures applications closely reflecting the condition of bed HydroQual, Inc., Mahwah, NJ, USA.
sediments prevailing in natural river channels. To better reflect Johnson PA (1992) Reliability-based pier scour engineering.

12
Water Management A bridge pier scour model with non-
uniform sediments
Pournazeri, Li and Haghighat

Journal of Hydraulic Engineering ASCE 118(10): 1344– Richardson JR and Wacker AM (1999) Bridge scour and stream
1358. instability counter measures in stream stability and scour
Khosronejad A, Kang S and Sotiropoulos F (2012) Experimental highway bridges. Compendium of Papers of ASCE Water
and computational investigation of local scour around bridge Resources Engineering Conferences 1991–1998 (Richardson
piers. Advances in Water Resources 37: 73–85. EV and Lagasse PF (eds)). ASCE, Reston, VA, USA, pp.
Kirkil G, Constantinescu G and Ettema R (2008) Coherent 939–946.
structures in the flow field around a circular cylinder with Roulund A, Sumer BM, Fredsoe J and Michelsen J (2005)
scour hole. Journal of Hydraulic Engineering ASCE 134(5): Numerical and experimental investigation of flow and scour
572–587. around a circular pile. Journal of Fluid Mechanics 534: 351–
Lee SO and Strum TW (2009) Effect of sediment size scaling on 401.
physical modeling of bridge pier scour. Journal of Hydraulic Sheppard D and Miller W (2006) Live-bed local pier scour
Engineering ASCE 135(10): 793–802. experiments. Journal of Hydraulic Engineering ASCE 132(7):
Liang D and Cheng L (2005) Numerical model for wave-induced 635–642.
scour below a submarine pipeline. Journal of Waterway, Port, Smagorinsky J (1963) General circulation experiments with the
Coastal, Ocean Engineering 131(5): 193–202. primitive equations. Monthly Weather Review 91(3): 99–164.
Mellor GL and Yamada T (1982) Development of a turbulence Tritthart M and Schober B (2011) Non-uniformity and layering in
closure model for geophysical fluid problems. Reviews of sediment transport modelling 1: flume simulations. Journal of
Geophysics and Space Physics 20(4): 851–875. Hydraulic Research 49(3): 325–334.
Melville BW and Chiew Y (1999) Time scale for local scour at van Rijn L (1984) Sediment transport, part I: bed load transport.
bridge piers. Journal of Hydraulic Engineering ASCE 125(1): Journal of Hydraulic Engineering ASCE 110(10): 1431–
59–54. 1456.
Melville B and Sutherland A (1988) Design method for local Wu W, Rodi W and Wenka T (2000) 3D numerical modeling of
scour at bridge piers. Journal of Hydraulic Engineering flow and sediment transport in open channels. Journal of
ASCE 114(10): 1210–1226. Hydraulic Engineering ASCE 126(1): 4–15.
Muzzammil M and Siddiqui NA (2009) A reliability-based Yang F, Liu X, Yang K and Cao S (2009) Study on the angle of
assessment of bridge pier scour in non-uniform sediments. repose of non-uniform sediment. Journal of Hydrodynamics
Journal of Hydraulic Research 47(3): 372–380. 21(5): 685–691.
Nagata N, Hosoda T, Nakato T and Muramoto Y (2005) Three-
dimensional numerical model for flow and bed deformation
around river hydraulic structures. Journal of Hydraulic
Engineering ASCE 131(12): 1074–1087.
Olsen NRB and Kjellesvig HM (1998) Three-dimensional
numerical flow modeling for estimation of maximum local
scour depth. Journal of Hydraulic Research 364(4): 579–590.
Olsen NRB and Melaaen MC (1993) Three-dimensional
calculation of scour around cylinders. Journal of Hydraulic
Engineering ASCE 119(9): 1048–1054.
Paik J, Escauriaza C and Sotiropoulos F (2010) Coherent structure
dynamics in turbulent flows past in-stream structures: some
insights gained via numerical simulation. Journal of
Hydraulic Engineering ASCE 136(12): 981–993.
Paphitis D (2001) Sediment movement under unidirectional
flows: an assessment of empirical threshold curves. Coastal WHAT DO YOU THINK?

Engineering 43(3–4): 227–245. To discuss this paper, please email up to 500 words to the
Parker G (1990) Surface-based bedload transport relation for editor at journals@ice.org.uk. Your contribution will be
gravel rivers. Journal of Hydraulic Research 28(4): 417–436. forwarded to the author(s) for a reply and, if considered
Pournazeri S, Li SS and Haghighat F (2013) An efficient multi- appropriate by the editorial panel, will be published as a
layer model for pier scour computations. Proceedings of the discussion in a future issue of the journal.
Institution of Civil Engineers – Water Management http:// Proceedings journals rely entirely on contributions sent in
dx.doi.org/10.1680/wama.13.00056. by civil engineering professionals, academics and students.
Raikar RV and Dey S (2005) Clear-water scour at bridge piers in Papers should be 2000–5000 words long (briefing papers
fine and medium gravel beds. Canadian Journal of Civil should be 1000–2000 words long), with adequate illustra-
Engineering 32(4): 775–781. tions and references. You can submit your paper online via
Raudkivi A (1986) Functional trends of scour at bridge piers. www.icevirtuallibrary.com/content/journals, where you
Journal of Hydraulic Engineering ASCE 112(1): 1–13. will also find detailed author guidelines.

13
View publication stats

Вам также может понравиться