You are on page 1of 7

International Journal of Sediment Research 37 (2022) 328e334

Contents lists available at ScienceDirect

International Journal of Sediment Research


journal homepage: www.elsevier.com/locate/ijsrc

Original Research

Collar performance in bridge pier scour with debris accumulation


Hossein Hamidifar a, *, Seyed Mohammad Bagher Shahabi-Haghighi a, Yee Meng Chiew b
a
Water Engineering Department, Shiraz University, Shiraz 71441, Iran
b
School of Civil and Environmental Engineering, Nanyang Technological University, Singapore 639798, Singapore

a r t i c l e i n f o a b s t r a c t

Article history: The mechanism of bridge pier scour becomes more complex in the presence of debris accumulation
Received 7 May 2021 upstream of the pier. While using countermeasures may be effective in reducing scour, their efficacy
Received in revised form could be undermined in such a situation. The current study investigates the effectiveness of using a collar
19 October 2021
in the presence of different types of floating debris accumulation in reducing scour around a cylindrical
Accepted 20 October 2021
Available online 29 October 2021
bridge pier with non-cohesive bed sediment. The experimental results reveal that using a collar can
reduce the scour depth by up to 39% when compared to that without the collar in the absence of debris
accumulation under the same flow conditions. Also, the collar's efficiency is reduced by up to 25% when
Keywords:
Bridge pier
debris accumulation is present. However, depending on the relative size of the accumulated debris and
Scour collar, the effectiveness of collars may be increased as a pier-scour protection device. A new parameter,
Debris which is defined as the Debris Protrusion Ratio (DPR), is used to quantify the effect of debris layout size
Collar on scour depth. Moreover, a semi-empirical equation that may be used to predict the depth of the scour
Blockage hole that forms around cylindrical collar-fitted piers with floating debris accumulation is proposed. The
equation's performance is evaluated using the available data in the literature. Based on the analysis, the
current study shows that the accuracy of the proposed equation is satisfactory in predicting the
maximum scour depth.
© 2021 International Research and Training Centre on Erosion and Sedimentation/the World Association
for Sedimentation and Erosion Research. Published by Elsevier B.V. All rights reserved.

1. Introduction resistance of the bed material against the flow. Some of the
methods that fall in this category are the use of riprap, gabions,
Scour at bridge piers represents a significant concern in hy- sand-filled geotextile bags, and grout-filled containers and mat-
draulic engineering. Factors such as the approach flow depth and tresses (Chiew, 1995; Chiew & Lim, 2000; Pagliara et al., 2010;
flow velocity (Chiew & Melville, 1987; Hamidifar et al., 2021; Yoon, 2005). In the second approach, devices are used to modify
Sheppard et al., 2004; Sheppard & Miller, 2006; Simarro et al., the flow field, thereby decreasing its erosive capacity. For
2007), sediment gradation and mixture (Chiew & Melville, 1989), example, streamlining the pier shape, placing a collar around the
angle of inclination of the pier relative to the flow (Yang et al., pier (Amini et al., 2017; Chiew, 1992; Heidarpour et al., 2010;
2020), pier shape, arrangement, foundation geometry (Keshavarzi Izadinia & Heidarpour, 2012; Karimaei Tabarestani & Zarrati,
et al., 2019, 2018; Panici & de Almeida, 2020a), and debris accu- 2019; Moncada-M et al., 2009; Valela et al., 2021; Zarrati et al.,
mulation (Ebrahimi et al., 2018; Panici & de Almeida, 2020b), affect 2004, 2010), implementing a slot in the pier (Chiew, 1992;
the development of the scour hole. Moncada-M et al., 2009), using a bed-sill in the vicinity of the pier
If the safety of a bridge is adversely affected by pier scour, (Gaudio et al., 2012; Grimaldi et al., 2009; Tafarojnoruz et al.,
countermeasures often are used to reduce the maximum scour 2012), installing sacrificial pile(s) upstream of the pier space
depth. Selection of the most suitable method for the protection of (Haque et al., 2007; Park et al., 2016a; Tafarojnoruz et al., 2012),
a bridge against scouring is based on the flow characteristics, site and utilizing tetrahedron frames and tethering a cable sur-
geometry, river morphology, and construction and maintenance rounding the pier (Izadinia & Heidarpour, 2012; Tang et al., 2009),
costs. Generally, two basic approaches are used in scour coun- are amongst some commonly used flow altering-
termeasures. The first approach deals with increasing the countermeasures for pier-scour reduction.
As a type of pier-scour countermeasure, a collar essentially is a
plate attached to the pier and installed at an elevation above, below,
* Corresponding author. or at the undisturbed river bed level. The efficiency of using collars
E-mail address: hamidifar@shitrazu.ac.ir (H. Hamidifar). in reducing the maximum scour depth under no-debris conditions

https://doi.org/10.1016/j.ijsrc.2021.10.002
1001-6279/© 2021 International Research and Training Centre on Erosion and Sedimentation/the World Association for Sedimentation and Erosion Research. Published by
Elsevier B.V. All rights reserved.
H. Hamidifar et al. / International Journal of Sediment Research 37 (2022) 328e334 329

has been investigated in previous studies (Chiew, 1992; Melville & University, Shiraz, Iran (Figs. 1(a)e1(e)). To reduce the formation of
Dongol, 1992; Zarrati et al., 2004). Recently, Valela et al. (2021) large-scale turbulence and surface waves at the flume entrance, a
introduced a new collar design and found that the new design flow straightener comprising an array of PVC pipes with 30 mm in
can reduce the maximum scour depth around the pier up to nearly diameter and 0.5 m in length were placed at the entrance. The
70% compared to the no-collar conditions. longitudinal slope of the flume was set at 0.002. The diameter and
Accumulation of floating debris around bridge piers during height of the bridge pier model (made of acrylic) were 40 mm and
floods is known to augment the scour hole, thereby potentially 500 mm, respectively. The bridge pier model was placed at a
triggering bridge failure (Lagasse et al., 2010). Generally, accu- location that was 5.5 m from the flume entrance. The thickness of
mulated debris with a particular shape form at the water surface the sediment bed into which the model pier was placed was
at the leading edge of the bridge pier (Diehl, 1997). This reduces 200 mm. The pier diameter (D) was chosen such that the effect of
the flow passage area, thereby increasing the risk of flooding in flume side walls is negligible as the ratio of the flume width (B) to
the neighboring regions. Additionally, floating materials can the pier diameter (B/D ¼ 10), which is customarily called the flume
destabilize the bridge deck and piers because of excessive hy- aspect ratio, exceeded 6.25 (Raudkivi & Ettema, 1983). To prevent
drodynamic loading (Oudenbroek et al., 2018). While in most ripple formation even in the clear-water condition, the median
cases the accumulated debris does not obstruct the entire flow particle size, d50, used in the current study, was 0.78 mm. Addi-
area, there are situations reported in the literature detailing how tionally, the geometrical standard deviation of the sediment par-
the debris has caused bridge overtopping and aggravated scouring ticles (sg ¼ d84/d16) was 1.29, where d84 and d16 are diameters for
(Diehl, 1997; Pagliara & Carnacina, 2013). In studying the effect of which 84% and 16%, respectively, of the sediment particles are finer.
debris accumulation on bridge piers and the related scour hole Flow discharges were measured using an electromagnetic
evolution, Pagliara and Carnacina (2010) found that the presence flowmeter with a precision of ±0.1 L/s installed at the inlet pipe. The
of debris can increase the maximum scour depth by up to three experiments were done under clear-water conditions since the
times compared to that without debris accumulation. Addition- flow intensity factor, i.e., the ratio of the undisturbed approach flow
ally, Park et al. (2016b) did a comprehensive set of experimental velocity (U) to critical velocity (Uc) for initiation of bed sediment
tests and found that the maximum scour depth around a single motion (U/Uc ¼ 0.83), was less than 1, which ensures that there is
pier with debris increased up to 60% compared to that without no transport of sediment into the scour hole from upstream. The
debris. Furthermore, Rahimi et al. (2018) did a set of experiments critical velocity was calculated according to Chiew and Melville
to study the effect of debris shape, thickness, length, and position (1987).
on the local scour at pier groups. They found that the maximum A preliminary experiment that lasted for 48 h was done with a
depth of the scour hole increases with the debris thickness and simple pier configuration to determine the time required for the
length. Moreover, Siddiq et al. (2019) found that when the debris rest of the study. The pier-scour depth at the flume centerline was
is attached to the pier at a depth of 0.75 relative to the water measured using a laser meter with ±0.1 mm resolution mounted on
surface height, it will cause more scouring compared to the pier a carriage. The measured temporal development of the scour depth
without debris and also when the debris is accumulated just was compared with the analytical approach proposed by Cheng
below the water surface. et al. (2016), showing that the 48 h duration was sufficient for the
While the effect of the hydraulic and geometrical factors on pier- scour hole to reach a quasi-equilibrium state. The results also
scour has been studied extensively, few studies have focused on the showed that approximately 90% of the final scour depth had
effect of debris accumulation. Because of the large variability of the occurred within the first 6 h. Hence, all subsequent experiments
position, size, shape, roughness, location, and permeability of the were done with a duration of 6 h. Nevertheless, this period is lower
accumulated debris, its effect on pier-scour is difficult to predict than that would be expected to achieve equilibrium scour depth
(Pagliara & Carnacina, 2010, 2011). Some researchers have intro- (Amini et al., 2012; Cheng et al., 2016). However, use of a 6 h
duced an equivalent diameter as the effective bridge pier diameter duration is supported by the study of Melville and Chiew (1999),
in the presence of debris (Lagasse et al., 2010; Melville & Dongol, who concluded that just 10% of the time needed to reach equilib-
1992). rium is required for the scour depth to attain 50%e80% of the
Although certain studies have been independently done on the maximum equilibrium scour depth. While a longer time is needed
effect of collars as well as debris accumulation on the bridge pier for the equilibrium scour conditions (e.g., the experiment done by
scour, a key question remains unanswered: how much does the Melville and Dongol (1992) lasted for more than 150 h), the 6 h
efficacy of using collars as a scour countermeasure change in the duration, which is the same or even longer than some similar
presence of debris accumulation upstream of a cylindrical bridge studies reported in the literature (Ebrahimi et al., 2018; Müller
pier? In an attempt to answer this question, the current study et al., 2001), was found to be sufficient for the scour hole to reach
investigates clear-water scour around a cylindrical bridge pier in a quasi-equilibrium condition in the current study. Moreover, time
the presence of a collar and debris accumulation with non- effects could be neglected in the current study because the aim is to
cohesive bed sediment. Moreover, the effect of the shape and compare the scour hole with and without debris accumulation after
size of the accumulated debris as well as the effects of flow depth the same time duration of testing.
on the efficiency of the collar also are examined. Finally, a semi- A metal plate of 100 mm in diameter (W ¼ 2.5D) and 1 mm in
empirical equation for the determination of the maximum scour thickness was fitted to the pier at the initial bed level elevation to
depth around circular piers in the presence of a collar and/or simulate the collar (see Fig. 1). The diameter (W) and position of the
debris is proposed and its performance is evaluated by compari- collar were selected based on the studies of Zarrati et al. (2004) and
son with some well-known equations proposed in previous Moncada-M et al. (2009).
studies. In order to simulate debris accumulation upstream of the pier,
six simplified debris shapes made of Teflon were used, namely two
2. Materials and methods cylindrical shapes, an inverted half-pyramid (triangular base), and
three rectangular plates (Fig. 1(e)), which are based on the studies
The experiments in this study were done in a recirculating glass- of Lagasse et al. (2010) and Ebrahimi et al. (2018). The cylindrical
walled rectangular flume that was 9 m long, 0.4 m wide, and 0.6 m shape represents a tree trunk, which is the most common shape of
high located at the Sediment Hydraulic Laboratory, Shiraz woody debris (Diehl, 1997) and gives the deepest scour based on
330 H. Hamidifar et al. / International Journal of Sediment Research 37 (2022) 328e334

Fig. 1. (a) Schematic of the experimental flume; (b) front view, (c) side view, (d) plan view of pier, debris, and collar, respectively; and (e) cylindrical, inverse pyramid, and
rectangular plate shapes of debris.

the experimental study of Melville and Dongol (1992). The For example, Test C-D1-130 signifies the test with a collar fitted to
inverted half-pyramid shape is the typical accumulation shape of the pier with a 12-mm-diameter cylindrical debris and 130 mm
a group of non-uniform size debris elements and has been used by flow depth.
previous researchers (Melville & Dongol, 1992). The rectangular
shape represents a mat of small woody elements. The dimensions 3. Results and discussion
of the debris shapes, which are listed in Table 1, are selected to be
within the range reported by Pagliara and Carnacina (2011). In To quantitatively evaluate how much the maximum scour depth
Table 1, Lx, Ly, and Lz are lengths in the streamwise, transverse, and changed after adding each combination of collar and debris,
vertical directions, respectively. In all the experiments, debris was changes of the maximum scour depth relative to the control test
located just under the free surface because the maximum scour (NC-ND test), Rnc, were calculated and listed in Table 2, in which Rnc
depth occurs when the debris is located there (Ebrahimi et al., is calculated as
2018).
Table 2 lists the 18 tests that were done in the current study. A dsc  dsn
Rnc ¼  100% (1)
unique code is assigned to each experiment. The first part of the dsn
code indicates whether a collar is present or not, with the letters C
and NC representing tests with and without collars, respectively. where dsc and dsn are the maximum scour depth around a collar-
The second part indicates the debris shape, with D1 and D2 fitted pier and simple pier without collar, respectively. The results
denoting that the debris is in the form of the 12- and 24-mm- show that the collar reduces the maximum scour depth by up to
diameter cylinders, respectively. D3 symbolizes the inverted pyra- 39.1% for the test without debris and up to 34%, 27%, 25%, 42%, 32%,
mid shape while D4, D5, and D6 represent rectangular shapes with and 29% for the D1, D2, D3, D4, D5, and D6 debris shapes, respec-
different sizes, as listed in Table 1. The third part indicates the tively, compared to the simple pier case.
magnitude of the undisturbed approach flow depth in millimeter. An attempt also has been made to evaluate how much the
maximum scour depth changes due to the presence of debris with
respect to the no-debris conditions (C-ND test). Hence, variations of
Table 1 the debris-enhanced maximum scour depth relative to the collar-
Characteristics of debris used in the experiments.
fitted pier without debris, Rc, was calculated and listed in Table 2,
Debris shape Symbol Lx (mm) Ly (mm) Lz (mm) in which Rc is calculated as
Cylinder D1 12 200 12
Cylinder D2 24 200 24 dsd  dsc
Rc ¼  100% (2)
Inverse pyramid D3 24 200 24 dsc
Rectangular plate D4 100 200 12
Rectangular plate D5 75 200 12 where dsd is the maximum scour depth around the bridge pier in
Rectangular plate D6 50 200 12
the tests with debris accumulation. It can be seen in Table 2 that the
H. Hamidifar et al. / International Journal of Sediment Research 37 (2022) 328e334 331

Table 2
Summary of the experimental conditions.

Run Code Scour countermeasure Debris type H (mm) Q (m3/s) Fr Rnc (%) Rc (%)

1 NC-ND-130 No-collar e 130 0.0129 0.22 e e


2 NC-ND-160 No-collar e 160 0.0153 0.19 e e
3 NC-ND-180 No-collar e 180 0.0167 0.17 e e
4 C-ND-130 Collar e 130 0.0129 0.22 39.13 e
5 C-ND-160 Collar e 160 0.0153 0.19 37.05 e
6 C-ND-180 Collar e 180 0.0167 0.17 36.86 e
7 C-D1-130 Collar D1 130 0.0129 0.22 32.66 10.64
8 C-D2-130 Collar D2 130 0.0129 0.22 27.48 19.15
9 C-D3-130 Collar D3 130 0.0129 0.22 24.89 23.40
10 C-D4-130 Collar D4 130 0.0129 0.22 40.43 2.13
11 C-D5-130 Collar D5 130 0.0129 0.22 32.79 10.42
12 C-D6-130 Collar D6 130 0.0129 0.22 29.51 15.81
13 C-D1-160 Collar D1 160 0.0153 0.19 34.38 4.26
14 C-D2-160 Collar D2 160 0.0153 0.19 26.34 17.02
15 C-D3-160 Collar D3 160 0.0153 0.19 25.00 19.15
16 C-D4-160 Collar D4 160 0.0153 0.19 39.73 4.26
17 C-D5-160 Collar D5 160 0.0153 0.19 28.57 13.48
18 C-D6-160 Collar D6 160 0.0153 0.19 21.43 24.82
19 C-D1-180 Collar D1 180 0.0167 0.17 34.12 4.35
20 C-D2-180 Collar D2 180 0.0167 0.17 27.25 15.22
21 C-D3-180 Collar D3 180 0.0167 0.17 24.51 19.57
22 C-D4-180 Collar D4 180 0.0167 0.17 42.35 8.70

dsc  dsn d  dsc


H: flow depth, Q: flow discharge, Fr: Froude number, Rnc ¼  100%, Rc ¼ sd  100%.
dsn dsc

D1, D2, D3, D5, and D6 debris types resulted in an increase of the
maximum scour depth up to 10%, 17%, 23%, 13%, and 24%, respec-
tively, while the D4 debris shape decreases the maximum scour
depth up to 8.7% relative to the collar-fitted pier without debris. The
data reveal that interactions of the collar and D4 debris are effective
in decreasing the maximum scour depth compared to the combi-
nation of collars with other debris types. One may infer from the
reduction observed for the D4-shape that the flow structure is
considerably affected by the debris type, and hence, the opposite
trend of Rc-values. Although the flow field was not measured in the
current study, Pagliara and Carnacina (2013) have observed that
due to the contraction caused by debris accumulation, the kinetic
energy of the flow at the bed around the pier increases, thereby
increasing the flow's ability to erode the bed sediment.
The flow section reduction, as was previously discussed, also
affects the vertical velocity. In the current study, visual observation
reveals that the debris size and shape can dictate how the flow is
diverted, i.e., directed either toward the bed or along the main flow Fig. 2. Variation of dimensionless maximum scour depth (dsd/dsn) as function of Debris
direction. The D1, D2, and D3 debris types have smaller lengths in Protrusion Ratio (DPR).
the streamwise direction (Lx) than the collar protrusion, i.e., Cp ¼
(WD)/2, as shown in Fig. 1(c). Consequently, a new parameter
called Debris Protrusion Ratio, DPR ¼ Lx/Cp, is proposed and its hence, causing a smaller scour depth. Changes in the flow direction
influence on the dimensionless maximum scour depth dsd/dsn (ratio are schematically illustrated in Fig. 3.
of the debris-accumulated to the no-debris conditions) is plotted in An attempt is made to derive a simple semi-empirical formula to
Fig. 2. Based on the experimental data, a regression curve, Eq. (3), determine the maximum scour depth in the presence of debris in
with a coefficient of determination, R2 ¼ 0.88, is fitted to the data. front of a collar-fitted pier. Based on the hypothesis of flow division
Eq. (3), which is fitted using data limited to the debris shapes and over and under the bridge deck proposed by Umbrell et al. (1998),
flow characteristics (e.g., clear-water scour conditions) listed in the total flow discharge (Q) passing a collar-fitted pier with floating
Tables 1 and 2, is as follows: debris accumulation may be written as
dsd
¼  0:099DPR2 þ 0:315DPR þ 1 (3) Q ¼ Qabove þ Qbetween þ Qbelow (4)
dsn
Fig. 2 shows that the maximum scour depth occurs at DPR z 1.6 where Qabove is the discharge above the debris; Qbelow is the
and the scour depth is lower for both smaller and larger values of discharge below the collar; and Qbetween is the discharge between
this DPR-value. Comparing the six debris types tested in the current the collar and the debris. As a first approximation, the following
study, it can be seen that for D1, D2, and D3 with DPR < 1.6, the flow assumptions are made in the analysis: (1) the velocity above the
is observed to have diverted toward the bed, which likely increased debris is equal to the undisturbed approach velocity of the flow (U);
the near-bed flow kinetic energy (Pagliara & Carnacina, 2013), and (2) the velocity below the debris is equal to the critical flow velocity
hence, greater scour depths are expected. However, D4 with DPR > for sediment entrainment (Uc), befitting the situation when the
1.6 seems to have redirected the flow directly on top of the collar, scour hole has reached its equilibrium condition.
332 H. Hamidifar et al. / International Journal of Sediment Research 37 (2022) 328e334

function of the flow depth to pier diameter ratio, incorporating the


mutual effects of collar and debris on the maximum scour depth.
The current experimental data are used to fit Eq. (7) to obtain
the form of f(H/D), resulting in the plot in Fig. 4.

H
f ðH = DÞ ¼ 0:19 þ 0:61 (8)
D
Eq. (8) is valid for 3.25 < H/D < 4.5, based solely on the current
test data. By substituting Eqs. (6) and (8) into Eq. (7), Eq. (9) which
may be used to compute the maximum scour depth in the presence
of both debris and collar, is obtained. It should be noted that Eq. (9)
is only valid for the experimental ranges used in the current study,
i.e., a collar, with diameter twice that of the pier, is placed at the
initial (undisturbed) bed level, and cylindrical, inverse pyramid,
and rectangular shaped debris located just below the free surface of
the approaching subcritical flow. Clearly, the accuracy of Eq. (9)
needs to be further tested with more experimental data with
different flow, debris, and sediment conditions. Using the current
data, the fitted equation is
 
ds H U Lz
¼  0:81 þ þ 0:61 (9)
D D Uc D
An attempt is made to compare the maximum scour depth
calculated by using Eq. (9) with published data and the results are
shown in Fig. 5, in which the subscripts o and c refer to the
observed and calculated values, respectively. It should be pointed
Fig. 3. Schematic illustration of the effect of debris size along the main flow direction out that none of the previous studies used for comparison in Fig. 5
on the flow path in the presence of a collar: (a) DPR z 1.6, and (b) DPR > 1.6 or DPR < has considered the combined effect of debris and a collar. Yet, each
1.6. of these studies deals with a collar or debris as listed in Table 3.
Moreover, the accuracy of the maximum scour depth calculated by
using Eq. (9) and those extracted from previous studies is evaluated
It should be noted that the aforementioned assumptions were using the Normalized Root Mean Square Error (NRMSE) index
used by Ebrahimi et al. (2018) in their study on a sharp-nose bridge which is defined as
pier with debris blockage. While these assumptions may not be rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P n 2
ðds;c ds;o Þ
strictly applicable to the zone between the debris and collar, the i¼1
n
findings of Ebrahimi et al. (2018) and those of the current study, NRMSE ¼ (10)
ds;u  ds;l
which are discussed later, yield satisfactory results. With these
assumptions, and dividing both sides of Eq. (4) by D, Eq. (4) may be where i is the counter; n is the total number of observed values;
re-written as follows for a unit width of the channel in the region ds,c and ds,o are the calculated and observed scour depths,
occupied by the scour hole: respectively; and ds,u and ds,l are the upper and lower limits of the
maximum observed scour depths, respectively. The calculated
UH UHa Uc ðH  Lz Þ Uc dsp values of NRMSE are listed in Table 3, which indicate that Eq. (9)
¼ þ þ (5)
D D D D has the best performance with a minimum value of
NRMSE ¼ 0.17. Amongst the methods developed for bridge pier
where Ha is the water depth above the debris and dsp is the
maximum equilibrium scour depth. In the current study where
debris is located just below the free surface (Ha ¼ 0) and the collar is
fitted to the pier at the initial bed level, Eq. (5) can be simplified as
 
dsp H U Lz
¼ 1 þ (6)
D D Uc D
Previous studies have shown that the maximum scour depth in
the presence of collar or debris is related to the approach flow
depth, H (Ebrahimi et al., 2018; Moncada-M et al., 2009), as is
implied in Eq. (6). Considering the assumptions made in deriving
Eq. (6), as well as the complex flow around the bridge pier in the
presence of both debris and collar, ds/D, the maximum scour depth
ratio may be simply written as

ds dsp
¼ þ f ðH = DÞ (7)
D D

where ds is the maximum scour depth at the end of the test with
any pier configurations other than the simple pier and f(H/D) is a Fig. 4. Variations of f(H/D) with H/D.
H. Hamidifar et al. / International Journal of Sediment Research 37 (2022) 328e334 333

Fig. 5. Calculated using Eq. (9) versus observed from literature and current studies maximum scour depth values.

Table 3
Comparison of the results of some previous studies with those of the present study*. countermeasure. The following conclusions are drawn from the
current study:
Reference Collar (W/D) Debris NRMSE

Tanaka and Yano (1967) þ (3) e 0.38 (1) The debris accumulated upstream of a pier just under the
Ettema (1980) þ (2) e 0.59
Chiew (1992) þ (2) e 0.53
free surface increases the maximum scour depth in most
Melville and Dongol (1992) e þ 3.57 cases, and there are situations when the mutual interac-
Zarrati et al. (2004) þ (2) e 0.62 tion of the collar and debris is more effective in protecting
Zarrati et al. (2004) þ (3) e 1.44 the pier against scouring than would result with collars
Pagliara and Carnacina (2010) e þ 4.55
alone.
Pagliara and Carnacina (2011) e þ 1.19
Jahangirzadeh et al. (2014) þ (2.5) e 0.27 (2) A new dimensionless parameter, namely the Debris Protru-
Ebrahimi et al. (2018) e þ 0.60 sion Ratio, DPR, is used to incorporate the inter-related effect
Eq. (9) þ (2.5) þ 0.17 of the collar and debris size. The value of 1.6 is found to be the
þ Presence of the collar or debris. critical DPR corresponding to the maximum scour depth, ds.
e Non-presence of the collar or debris. (3) An empirical equation is fitted to the experimental data of
* Values in the parenthesis are the ratio of the collar diameter to the pier diameter the current study to relate the maximum scour depth as a
(W/D).
function of DPR.
(4) Based on the theory of mass flow conservation, a semi-
scour computation in the presence of debris, the method of empirical equation for the determination of the maximum
Ebrahimi et al. (2018) has an acceptable performance with an scour depth around circular piers with a collar and debris
NRMSE ¼ 0.60. Moreover, the methods developed for collar- accumulation is proposed. When compared with published
fitted piers, namely that of Jahangirzadeh et al. (2014), Tanaka studies on scour with a collar or debris, the proposed
and Yano (1967), Chiew (1992), and Ettema (1980) show satis- equation gives reasonable results based on the NRMSE
factory results with NRMSE values of 0.27, 0.38, 0.53, and 0.59, parameter.
respectively; although none of them considered debris in their (5) The results of the current study may be used to estimate the
studies. Additionally, the NRMSE values obtained using the maximum scour depth around circular bridge piers in the
Pagliara and Carnacina (2010) and Melville and Dongol (1992) presence of both a collar and debris in preliminary designs.
methods are the highest among all of the methods considered More studies with different experimental set-ups under
in Table 3. The considerable deviations between the results of different flow, debris, collar, and sediment conditions should
the current study and some of these studies such as the Pagliara be done to generalize the findings.
and Carnacina (2010) and Melville and Dongol (1992) methods Research data
do not imply inefficiency of these methods and may be attrib-
uted to the uncertainties arising from the different type, size, Data will be available from the corresponding author upon
and position of the debris. These discrepancies suggest that reasonable request.
more research is needed with various sizes and types of debris
under different hydraulic conditions, pier geometry and Funding
different scour countermeasures for a general scour prediction
model. This research did not receive any specific grant from funding
agencies in the public, commercial, or not-for-profit sectors.
4. Conclusions
Declaration of competing interest
In the current study, six different types of debris accumulation,
namely two cylindrical, an inverse half-pyramid, and a rectangular The authors declare that they have no known competing
plate, are used to examine the effect of the size and shape of debris financial interests or personal relationships that could have
on the efficiency of a collar fitted to a circular bridge pier as a scour appeared to influence the work reported in this paper.
334 H. Hamidifar et al. / International Journal of Sediment Research 37 (2022) 328e334

References Müller, G., Mach, R., & Kauppert, K. (2001). Mapping of bridge pier scour with
projection moire . Journal of Hydraulic Research, 39(5), 531e537.
Oudenbroek, K., Naderi, N., Bricker, J., Yang, Y., van der Veen, C., Uijttewaal, W.,
Amini, N., Balouchi, B., & Shafai Bejestan, M. (2017). Reduction of local scour at river
Moriguchi, S., & Jonkman, S. (2018). Hydrodynamic and debris-damming failure
confluences using a collar. International Journal of Sediment Research, 32(3),
of bridge decks and piers in steady flow. Geosciences, 8(11), 409.
364e372.
Pagliara, S., & Carnacina, I. (2010). Temporal scour evolution at bridge piers: Effect
Amini, A., Melville, B. W., Ali, T. M., & Ghazali, A. H. (2012). Clear-water local scour
of wood debris roughness and porosity. Journal of Hydraulic Research, 48(1),
around pile groups in shallow-water flow. Journal of Hydraulic Engineering,
3e13.
138(2), 177e185.
Pagliara, S., & Carnacina, I. (2011). Influence of wood debris accumulation on bridge
Cheng, N. S., Chiew, Y. M., & Chen, X. (2016). Scaling analysis of pier-scouring
pier scour. Journal of Hydraulic Engineering, 137(2), 254e261.
processes. Journal of Engineering Mechanics, 142(8), 6016005.
Pagliara, S., & Carnacina, I. (2013). Bridge pier flow field in the presence of debris
Chiew, Y. M. (1992). Scour protection at bridge piers. Journal of Hydraulic Engi-
accumulation. Proceedings of the Institution of Civil Engineers - Water Manage-
neering, 118(9), 1260e1269.
ment, 166(4), 187e198.
Chiew, Y. M. (1995). Mechanics of riprap failure at bridge piers. Journal of Hydraulic
Pagliara, S., Carnacina, I., & Cigni, F. (2010). Sills and gabions as countermeasures at
Engineering, 121(9), 635e643.
bridge pier in presence of debris accumulations. Journal of Hydraulic Research,
Chiew, Y. M., & Lim, F.-H. (2000). Failure behavior of riprap layer at bridge piers
48(6), 764e774.
under live-bed conditions. Journal of Hydraulic Engineering, 126(1), 43e55.
Panici, D., & de Almeida, G. A. M. (2020a). A theoretical analysis of the fluid-solid
Chiew, Y. M., & Melville, B. W. (1987). Local scour around bridge piers. Journal of
interactions governing the removal of woody debris jams from cylindrical
Hydraulic Research, 25(1), 15e26.
bridge piers. Journal of Fluid Mechanics, 886, https://doi.org/10.1017/
Chiew, Y. M., & Melville, B. W. (1989). Local scour at bridge piers with non-uniform
jfm.2019.1048
sediments. Proceedings - Institution of Civil Engineers, 87(2), 215e224.
Panici, D., & de Almeida, G. A. M. (2020b). Influence of pier geometry and debris
Diehl, T. H. (1997). Potential drift accumulation at bridges. Publication No. FHWA-RD-
characteristics on wood debris accumulations at bridge piers. Journal of Hy-
97-028. V.A., USA: U.S. Department of Transportation, Federal Highway
draulic Engineering, 146(6), 04020041.
Administration, Mc Lean.
Park, J. H., Sok, C., Park, C. K., & Kim, Y. D. (2016a). A study on the effects of debris
Ebrahimi, M., Kripakaran, P., Prodanovi c, D. M., Kahraman, R., Riella, M., Tabor, G.,
accumulation at sacrificial piles on bridge pier scour: I. Experimental results.
Arthur, S., & Djordjevi c, S. (2018). Experimental study on scour at a sharp-nose
KSCE Journal of Civil Engineering, 20(4), 1546e1551.
bridge pier with debris blockage. Journal of Hydraulic Engineering, 144(12),
Park, J. H., Sok, C., Park, C. K., & Kim, Y. D. (2016b). A study on the effects of debris
4018071.
accumulation at sacrificial piles on bridge pier scour: II. Empirical formula. KSCE
Ettema, R. (1980). Scour at bridge piers. Auckland, New Zealand: University of
Journal of Civil Engineering, 20(4), 1552e1557.
Auckland.
Rahimi, E., Qaderi, K., Rahimpour, M., & Ahmadi, M. M. (2018). Effect of debris on
Gaudio, R., Tafarojnoruz, A., & Calomino, F. (2012). Combined flow-altering coun-
piers group scour: An experimental study. KSCE Journal of Civil Engineering,
termeasures against bridge pier scour. Journal of Hydraulic Research, 50(1),
22(4), 1496e1505.
35e43.
Raudkivi, A. J., & Ettema, R. (1983). Clear-water scour at cylindrical piers. Journal of
Grimaldi, C., Gaudio, R., Calomino, F., & Cardoso, A. H. (2009). Countermeasures
Hydraulic Engineering, 109(3), 338e350.
against local scouring at bridge piers: Slot and combined system of slot and bed
Sheppard, D. M., & Miller, W. (2006). Live-bed local pier scour experiments. Journal
sill. Journal of Hydraulic Engineering, 135(5), 425e431.
of Hydraulic Engineering, 132(7), 635e642.
Hamidifar, H., Zanganeh-Inaloo, F., & Carnacina, I. (2021). Hybrid scour depth pre-
Sheppard, D. M., Odeh, M., & Glasser, T. (2004). Large scale clear-water local pier
diction equations for reliable design of bridge piers. Water, 13(15), 2019.
scour experiments. Journal of Hydraulic Engineering, 130(10), 957e963.
Haque, M. A., Rahman, M. M., Islam, G. T., & Hussain, M. A. (2007). Scour mitigation
Siddiq, A., Pasha, G., Ghani, U., & Ahmed, A. (2019). Impact of large wood debris
at bridge piers using sacrificial piles. Journal of Sedimentary Research, 22(1),
(LWD) accumulation on scour characteristic at bridge pier. Proceedings of the
49e59.
2nd International Conference on Sustainable Development in Civil Engineering,
Heidarpour, M., Afzalimehr, H., & Izadinia, E. (2010). Reduction of local scour around
545e548.
bridge pier groups using collars. International Journal of Sediment Research,
Simarro, G., Teixeira, L., & Cardoso, A. H. (2007). Flow intensity parameter in pier
25(4), 411e422.
scour experiments. Journal of Hydraulic Engineering, 133(11), 1261e1264.
Izadinia, E., & Heidarpour, M. (2012). Simultaneous use of cable and collar to pre-
Tafarojnoruz, A., Gaudio, R., & Calomino, F. (2012). Evaluation of flow-altering
vent local scouring around bridge pier. International Journal of Sediment
countermeasures against bridge pier scour. Journal of Hydraulic Engineering,
Research, 27(3), 394e401.
138(3), 297e305.
Jahangirzadeh, A., Basser, H., Akib, S., Karami, H., Naji, S., & Shamshirband, S. (2014).
Tanaka, S., & Yano, M. (1967). Local scour around a circular cylinder. Proceedings of
Experimental and numerical investigation of the effect of different shapes of
the 12th Congress of the International Association for Hydraulic Research,
collars on the reduction of scour around a single bridge pier. PLoS One, 9(6),
193e201.
98592.
Tang, H. W., Ding, B., Chiew, Y. M., & Fang, S. L. (2009). Protection of bridge piers
Karimaei Tabarestani, M., & Zarrati, A. R. (2019). Local scour depth at a bridge pier
against scouring with tetrahedral frames. International Journal of Sediment
protected by a collar in steady and unsteady flow. Proceedings of the Institution
Research, 24(4), 385e399.
of Civil Engineers - Water Management, 172(6), 301e311.
Umbrell, E. R., Young, G. K., Stein, S. M., & Jones, J. S. (1998). Clear-water contraction
Keshavarzi, A., Hamidifar, H., & Khajenoori, L. (2019). Mean flow structure and local
scour under bridges in pressure flow. Journal of Hydraulic Engineering, 124(2),
scour around single and two columns bridge piers. Irrigation Science and En-
236e240.
gineering, 42(4), 75e90.
Valela, C., Rennie, C. D., & Nistor, I. (2022). Improved bridge pier collar for reducing
Keshavarzi, A., Shrestha, C. K., Zahedani, M. R., Ball, J., & Khabbaz, H. (2018). Exper-
scour. International Journal of Sediment Research, 37(1), 37e46.
imental study of flow structure around two in-line bridge piers. Proceedings of
Yang, Y., Melville, B. W., Macky, G. H., & Shamseldin, A. Y. (2020). Temporal evo-
the Institution of Civil Engineers - Water Management, 171(6), 311e327.
lution of clear-water local scour at aligned and skewed complex bridge piers.
Lagasse, P. F., Zevenbergen, L. W., & Clopper, P. E. (2010). Impacts of debris on bridge
Journal of Hydraulic Engineering, 146(4), 04020026.
pier scour. Scour and Erosion, 210, 854e863.
Yoon, T. H. (2005). Wire gabion for protecting bridge piers. Journal of Hydraulic
Melville, B. W., & Chiew, Y. M. (1999). Time scale for local scour at bridge piers.
Engineering, 131(11), 942e949.
Journal of Hydraulic Engineering, 125(1), 59e65.
Zarrati, A. R., Chamani, M. R., Shafaie, A., & Latifi, M. (2010). Scour countermeasures
Melville, B. W., & Dongol, D. M. (1992). Bridge pier scour with debris accumulation.
for cylindrical piers using riprap and combination of collar and riprap. Inter-
Journal of Hydraulic Engineering, 118(9), 1306e1310.
national Journal of Sediment Research, 25(3), 313e322.
Moncada-M, A. T., Aguirre-Pe, J., Bolívar, J. C., & Flores, E. J. (2009). Scour protection
Zarrati, A. R., Gholami, H., & Mashahir, M. B. (2004). Application of collar to control
of circular bridge piers with collars and slots. Journal of Hydraulic Research,
scouring around rectangular bridge piers. Journal of Hydraulic Research, 42(1),
47(1), 119e126.
97e103.

You might also like