You are on page 1of 12

WATER RESOURCES RESEARCH, VOL. 42, W02405, doi:10.

1029/2005WR004139, 2006

Local scouring due to turbulent water jets downstream of a


trapezoidal drop: Laboratory experiments and stability analysis
Claudia Adduce and Michele La Rocca
Dipartimento di Scienze dellIngegneria Civile, Universita` degli Studi Roma Tre, Rome, Italy
Received 25 March 2005; revised 29 July 2005; accepted 8 November 2005; published 16 February 2006.

[1] Three series of laboratory experiments were run, varying the flume water discharge

and the downstream water depth, to investigate local scouring downstream of a


trapezoidal drop followed by a rigid apron. Three different kinds of jets, developing
downstream of the drop, were observed: a submerged jet, a surface wave jet, and an
oscillating jet. The oscillating jet is defined as a jet which changes intermittently between
a surface wave jet and a submerged jet. A different development of the scouring
process due to the different kinds of jet was observed: a submerged jet produced a local
scouring with a downstream dune, while a surface wave jet produced a local scouring
with both an upstream and a downstream dune; the scouring process due to the oscillating
jet was the result of the interaction of the scouring processes due to a submerged jet
and a surface wave jet. A stability analysis of the surface wave jet was performed, with
both a flat and a scoured bed, by means of the modified De SaintVenant equations,
that is, with correction terms accounting for the curvature of the streamlines. The stability
of the surface wave jet is affected by the presence of the scoured bed, which weakens its
stability properties.
Citation: Adduce, C., and M. La Rocca (2006), Local scouring due to turbulent water jets downstream of a trapezoidal drop:
Laboratory experiments and stability analysis, Water Resour. Res., 42, W02405, doi:10.1029/2005WR004139.

1. Introduction
[2] Drop structures are used to prevent excessive channel
bed degradation in rivers [Wu and Rajaratnam, 1998], but
the erosive action of the water flowing over it causes
downstream local scouring. The flow at the edge of a drop
is accelerated by gravity and may become supercritical,
behaving as an overfall jet. Jet flow due to the interaction of
the mainstream with hydraulic structures, such as drops,
gates, and spillways, was observed by several studies,
performed with a flat rigid bed or a mobile bed [Adduce
and Mele, 2004; Chatterjee et al., 1994; Dey and Westrich,
2003; Farhoudi and Smith, 1985; Gijs and Hoffmans, 1998;
Hassan and Narayanan, 1985; Kurniawan et al., 2001; Wu
and Rajaratnam, 1998; Fritz and Hager, 1998; Ohtsu and
Yasuda, 1991b]. When the jet impinges on a mobile bed, it
lifts the sediments, which are transported downstream, and a
scour hole is formed. Scouring can be so deep as to
endanger the stability of the structure and cause a risk of
failure if the foundations are not designed taking into
account the maximum scour depth.
[3] Most of the previous investigations on local scouring
were focused on developing empirical relations for the
maximum scour depth and length for noncohesive
[Schoklitsch, 1932; Bormann and Julien, 1991; Farhoudi
and Smith, 1985; Breusers and Raudkivi, 1991; Hoffmans,
1997] and cohesive [Dey and Westrich, 2003; Mazurek et al.,
2003] soils A comparative review of the existing formulae
can be found in work by Hoffmans and Verheij [1997].
Copyright 2006 by the American Geophysical Union.
0043-1397/06/2005WR004139

[4] The water flowing downstream of a hydraulic structure


may generate different kinds of flow, related both to different
hydraulic conditions and to hydraulic structures geometry
[see Ohtsu and Yasuda, 1991a, 1991b; Hager, 1992; Wu and
Rajaratnam, 1998; Fritz and Hager, 1998]. In particular, for
submerged overflow structures, with the downstream water
depth higher than the structures height, three flow conditions may occur, depending on the downstream water depth,
in order of rising downstream water depth: (1) plunging
flow, also referred to impinging flow, with a submerged main
flow and a surface roller; (2) surface wave (breaking or not
breaking) flow, with the main flow along the free surface and
a bottom recirculation zone; and (3) surface jet flow, similar
to the surface wave flow, but with a nearly horizontal free
surface [see Wu and Rajaratnam, 1998; Fritz and Hager,
1998; Bukreev, 2001]. Bukreev [2001] observed a transition
condition between the surface state (surface wave flow)
and the bottom state (plunging flow).
[5] Most of the previous studies investigated local scouring due to submerged jets developing downstream of a
hydraulic structure. A few studies investigated local scouring downstream of surface wave jets [Adduce, 2004; Gijs
and Hoffmans, 1998], so local scouring due to a surface
wave flow is still poorly understood. When a surface wave
jet develops downstream of a hydraulic structure, the local
scouring process may be very different from the case of a
scour hole produced by a submerged jet, because a different
kind of velocity field develops [Adduce, 2004]. In addition,
no theoretical study has been performed yet to investigate
the stability of the surface wave flow concerning the
transition between a surface wave jet and a submerged jet
found by Bukreev [2001].

W02405

1 of 12

W02405

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

[6] The aim of the present paper is to investigate local


scouring due to both submerged and surface wave jets, that
is, plunging and surface wave flow, respectively. Eight
scouring tests were run, changing both the water discharge
and the downstream water depth, in order to develop both
submerged jet flow (test H1 and H2) and surface wave flow
(tests U1, U2, and U3). An oscillating jet, that is, a transition
condition, which intermittently changes between a surface
wave jet and a submerged jet, was observed (tests T1, T2,
and T3). A different development of the scouring process
due to the different kinds of jets was observed and is
described in the paper. The velocity fields, measured in
the scour hole by an ultrasonic Doppler velocity profiler for
both a submerged jet and a surface wave jet, are shown. In
addition, the six experiments, with surface wave jets and
oscillating jets, were also performed with a flat bottom (tests
U1FB, U2FB, U3FB, T1FB, T2FB, and T3FB) in order to
highlight the influence of the scoured bed on the surface
wave jet stability. This latter has been theoretically investigated, starting from the analytical determination of the
velocity and pressure fields in presence of a surface wave
flow. Such velocity and pressure fields, unlike the work of
Montes and Chanson [1998], were obtained accounting for
a scoured bed. The velocity profiles were compared with the
experimental ones in order to have an idea of the validity of
the theoretical model. Velocity and pressure profiles were
used to perform integration of motion equations along the
vertical coordinate in order to obtain a modified version of
the De Saint Venant equations, that is, with correction
terms accounting for the curvature of the streamlines. A
stability analysis of the surface wave jet was performed, by
means of this modified version of De Saint Venant equations, with both a flat and a scoured bed. To the authors
knowledge, the stability analysis of a surface wave flow on
a scoured bed has never been performed before. Interesting
results are highlighted by the comparison between the
analysis on the surface wave flow on flat and scoured
bed: The scoured bed weakens the stability properties of
the surface wave flow.
[7] The paper is set out as follows: The experimental
apparatus is described in section 2; the laboratory measured
velocity fields and the scouring process are presented in
section 3; and the stability analysis of the surface wave jet
with both a flat and a scoured bed is described in section 4.
The conclusions of the work are presented in section 5.

2. Experimental Setup
[8] Laboratory experiments were conducted at the Hydraulics Laboratory of RomaTRE University, in a 17 m long,
1 m high, and 0.8 m wide flume of rectangular cross section.
Scouring tests were carried out in a 0.3 m high, 0.8 m wide
and L = 3 m long sediment recess section, positioned 7 m
downstream of the flume inlet and made by artificially
raising the flume bed. A uniformly graded sand characterized by d50 = 0.7 mm and a density rs = 2650 kg/m3 was
used to fill the sediment recess section in order to have
a mobile bed.
geometric standard deviation of the
qThe

d84
sand sg =
d16 = 1.21 was smaller than the threshold
proposed by Breusers and Raudkivi [1991] for the definition
of nonuniform grading (sg = 1.35). The same sand was
glued both to the upstream and downstream fixed-bed

W02405

Figure 1. Definition sketch of the test section in the x-z


plane.
sections to produce a bed of homogeneous roughness. All
the tests were conducted in clear water scour conditions. A
trapezoidal drop of height Dp = 0.15 m followed by a rigid
apron of length Lp = 0.5 m was positioned upstream of the
test section, as shown in Figure 1. A control gate, positioned
at the end of the flume, was used to change the downstream
water depth. The water level in the flume was measured
using a resistive point gauge of accuracy 0.10 mm,
mounted on a carriage moving along the flume. The water
discharge at the inlet was controlled by an inlet valve and
was measured by an electromagnetic flowmeter. For a
detailed description of the experimental apparatus, see
Adduce [2004].
[9] Before starting the experiments, the level of the
mobile sand bed, positioned in the sediment trap, was set
to the same height of the fixed bed. In order to avoid the
undesirable scour of the sediment bed, the control gate was
completely closed, and the flume was slowly filled with the
water pumped from the laboratory water reservoir by an
auxiliary pumping device. The water discharge was then
slowly increased, starting from zero, and was set to the
desired experimental value, paying attention to avoid any
movement of the sand bed. The experiment started when the
downstream water depth was set to the desired experimental
value by opening the control gate. The water discharge was
kept constant for the entire duration of the experiment.
Three series of scouring experiments were run, as shown in
Table 1: series H (submerged jet), series U (surface wave
jet), and series T (oscillating jet). We stopped the experiments when the measured maximum scour depth, plotted
versus logarithm of time, was sensibly constant [Adduce et
al., 2004; Adduce, 2004]. For the present experiments, the
maximum scour depth did not increase with the logarithm of
time after about 6 hours; as a consequence, the duration of
the experiments was set to 10 hours. The evolution of the
scour hole was recorded by a CCD camera connected to a
digital videocassette recorder. At the end of the experiment,
velocity measurements were performed by an ultrasonic
Doppler velocity profiler (UDVP).
[10] The ultrasonic Doppler velocity profiler, first developed for medical studies on blood flow Willemetz [1990], is
today used to perform measurements in both transparent [Ito
et al., 2001; Hurther, 2001; Alfonsi, 2001; Rolland, 1994]
and optically opaque fluids and high-temperature fluids
[Eckert and Gerberth, 2002], giving spatiotemporal velocity
information. The UDVP consists of a piezoelectric probe,
which works both as emitter and receiver of a signal
consisting of ultrasonic bursts traveling across the fluid

2 of 12

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

W02405

Table 1. Experimental Parameters: Discharge Q, Water Depth


Over the Drop hs, Downstream Water Depth Over the Rigid
Bottom h0, and Maximum Scour Depth zmax
Test

Q, L/s

hs, cm

h0, cm

zmax, m

H1
H2
T1 T1FB
T2 T2FB
T3 T3FB
U1 U1FB
U2 U2FB
U3 U3FB

31.20
36.45
30.00
34.04
38.64
31.01
33.10
37.40

3.80
4.10
4.20
4.45
4.80
4.80
5.00
5.20

16.19
16.91
18.57
18.86
19.34
19.88
20.36
20.78

0.107
0.139
0.040
0.056
0.108
0.051
0.055
0.062

and backscattered by the targets, that is, small air bubbles or


particles, moving with the flow. For a detailed description of
the UDVP technique, see Willemetz [1990, 2001], Rolland
[1994], and Takeda [1995, 1999]. The UDVP allows a
quasi-instantaneous measurement of the velocity in different
volumes (gates), positioned along the axis of the probe. The
distance between the centers of these gates used for the
experiments is 3.52 mm. The longitudinal size used for
the sampling volume is 2.96 mm, while its lateral size
depends on probe characteristics (emitting frequency and
diameter) and increases far from the probe [see Willemetz,
2001] because of the lateral spreading of the pressure wave.
During the present experiments, probes with an emitting
frequency of 2 MHz, a diameter of 14 mm, and a spreading
angle of 1.83 were used. The UDVP probe measures
profiles of the velocity component parallel to the probe
axis. In order to measure the profiles of the horizontal and
the vertical velocity components, three probes with the same
emitting frequency were used, following the technique
shown by Longo et al. [2001] and Adduce [2004]. Two
probes were tilted with respect to the vertical direction of an
angle of 20, while a third probe was positioned parallel to
the vertical direction. Velocity measurements were performed along the flume centerline. The considered values
of the experimental parameters, water discharge (Q), water
depth over the drop (hs), water depth downstream of the
sediment trap (h0), and maximum scour depth measured
along the flume centerline (zmax), are shown in Table 1.

3. Experimental Results
[11] Three series of scouring laboratory experiments
were run, corresponding to three different kinds of water
jets developing downstream of the drop: a submerged jet

Figure 2. Scour surface measured at the end of test H1.

W02405

(tests H1 and H2), a surface wave jet (tests U1, U2, and
U3), and an oscillating jet (tests T1, T2, and T3), related to
the development of a plunging flow, a surface wave flow,
and a transition flow, respectively. The oscillating jet is
defined as a jet which evolves intermittently between a
surface wave jet and a submerged jet, corresponding to the
intermittent evolution of a plunging flow into a surface
wave flow. A different behavior, concerning both the
development of the scour hole and the measured velocity
field, was observed for the different kinds of jets. In
addition, tests U1, U2, and U3 and T1, T2, and T3 were
also performed with a flat bottom in order to highlight the
influence of the scoured bed on the stability of the surface
wave jet. Such tests are called U1FB, U2FB, U3FB, T1FB,
T2FB, and T3FB, respectively, and their experimental
parameters are shown in Table 1.
3.1. Scouring Development
[12] When a plunging flow formed downstream of the
drop (tests H1 and H2), a submerged jet with a reverse flow
close to the water surface was observed [see Wu and
Rajaratnam, 1998; Fritz and Hager, 1998; Bukreev,
2001]. The scouring process developed very quickly during
the first minutes of the run, when both suspended load
transport and bed load transport were observed, and became
slower and slower as time advanced. A similar scouring
behavior was observed by Rajaratnam [1981], Farhoudi
and Smith [1985], and Adduce et al. [2004]. The sand
positioned close to the apron was moved downstream by
the submerged jet, and a downstream dune was formed. The
scour hole increased both in depth and length together with
the dune, which continued its downstream movement and
increased in length and height. At the end of the test, after
about 10 hours, measurements of the scour hole surface
were performed by a point gauge of accuracy 0.10 mm.
Figure 2 shows the measured scour surface due to a
submerged jet; a scour hole followed by a downstream
dune is observed. The scour hole and the downstream dune
are two dimensional.
[13] When a surface wave flow formed over the rigid
apron downstream of the drop (tests U1, U2, and U3), a
surface wave jet with a reverse flow close to the bottom was
observed [see Wu and Rajaratnam, 1998; Fritz and Hager,
1998]. The sand, positioned close to the rigid apron, was
moved upstream over the apron, creating a small dune,
which moved toward the drop. Far from the apron, the sand

Figure 3. Scour surface measured at the end of test U1.

3 of 12

W02405

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

W02405

was slowly moved downstream by the flow. At the end of


the test, two dunes were observed: the first one covering all
the apron and part of the drop, and the second one
positioned downstream over the mobile bed, as shown in
Figure 3. The dune over the apron was slightly three
dimensional, together with the scour hole and the downstream dune, which were symmetrical with respect to an
axis passing through the flume centerline.
[14] In tests T1, T2, and T3, a transitional condition was
observed: a flow which changed in time between a plunging
flow and a surface wave flow. In these tests, the velocity
field continued to change in time, showing the presence of
both a submerged jet and a surface wave jet. When a surface
wave jet was present (Figure 4a), the sand was partly moved
upstream over the apron and partly moved downstream over
the mobile bed, as for tests U1, U2, and U3. During the run,
the sand moving upstream formed a dune over the rigid
apron, with increasing height and length. Suddenly, the
surface wave flow developed into a plunging flow, with a
submerged jet, pushing fast downstream the sand present
over the apron (Figure 4b). The sand created a new dune,
which moved over the eroded bed and partly filled the scour
hole, as shown in Figures 4b and 4c. During the transition,
the observed velocity field was three dimensional, because
the surface wave jet and the submerged jet could be present
over the apron simultaneously in different longitudinal
sections of the flume (a longitudinal section is defined as
a vertical plane parallel to the flow direction). Therefore the
sand over the apron was moved partly downstream and
partly upstream, at the same time but in different longitudinal sections. The scour hole at the end of the test was
slightly three dimensional because of the presence of this
intermittent flow field.

Figure 4. Scouring temporal evolution for test T2 (a) 4 min


29 s after the beginning of the run, (b) 7 min 10 s after the
beginning of the run, (c) 7 min 41 s after the beginning of the
run, and (d) 23 min 3 s after the beginning of the run.

3.2. Velocity Fields


[15] The velocity field, the bed, and the free surface
profiles were measured along the flume centerline at the
end of the run. Figures 5 and 6 show the measured velocity
fields for tests H1 and U1, respectively. Because of instrument configuration, the data could not be obtained in
regions closer to 3.5 cm from the water surface and 1 cm
from the eroded bed. For the same reason, it was not
possible to perform velocity measurements in a zone approximately positioned over the rigid apron.
[16] Figure 5 shows the velocity field inside the scour
hole for test H1. Eleven velocity profiles, positioned at x =
5, 5, 15, 25, 35, 45, 55, 65, 75, 85, and 95 cm, following
the reference system shown in Figure 1, were measured.
The first profile at x = 5 cm shows a wall jet velocity
distribution along the vertical coordinate; that is, a hydraulic
jump is present over the apron downstream of the drop. A
similar velocity distribution, due to a hydraulic jump, was
observed by Rajaratnam [1965, 1976], Rajaratnam and
Murahari [1974], Hassan and Narayanan [1985], Hager
[1992], Kawagoshi and Hager [1990], and Fritz and Hager
[1998]. The wall jet dissipates quickly, as in the work by
Hassan and Narayanan [1985], because it produced a local
scour. A reverse flow at the top of the first measured profile
is shown. The wall jet velocity distribution develops into a
free jet pattern downstream of the apron, over the mobile
bed, as shown by the velocity profiles positioned at x = 5 cm
and x = 15 cm [see Adduce and Mele, 2004]. A weak reverse
flow close to the water surface is observed at x = 5 cm.

4 of 12

W02405

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

Figure 5. Velocity profiles of the vertical and horizontal


components, scour profile, and water surface profile
measured at the end of test H1.
Downstream of x = 15 cm, a small positive horizontal
component can be observed at the top of each velocity
profile. A small negative horizontal velocity component,
close to the bottom, is present for x = 15 cm, x = 25 cm, x =
35, and x = 45 cm. The negative velocities in run H1 show
the presence of a recirculation cell, due to the development
of a wall jet into a free jet. At the beginning of the
experiments, a submerged jet close to the bottom produced
a downstream sediment transport. As the time increased, a
scour hole and a recirculation cell gradually developed. This
recirculation cell did not produce any upstream sediment
transport, because of very small velocity values. From x =
55 cm to x = 95 cm, a positive horizontal velocity component close to both the top and the bottom of these velocity
profiles is observed.
[17] Figure 6 shows the velocity field inside the scour
hole of test U1. Nineteen velocity profiles, positioned at x =
5, 15, 25, 35, 45, 55, 65, 75, 85, 95, 105, 115, 125, 135,
145, 155, 165, 175, and 185, were measured following the

Figure 6. Velocity profiles of the vertical and horizontal


components, scour profile, and water surface profile
measured at the end of test U1, together with theoretical
scour profile (dashed lines) for test U1.

W02405

reference system shown in Figure 1. In this test, a surface


wave flow, with the main flow along the free surface and a
bottom recirculation zone, developed over the rigid apron,
that is, a surface wave jet [see Wu and Rajaratnam, 1998;
Fritz and Hager, 1998]. From x = 5 cm to about x = 75 cm,
the measured profiles show a large positive horizontal
velocity component at the top of the flow; that is, the
surface wave jet is not completely dissipated. A small
reverse flow close to the bottom can be observed at x =
5 cm, x = 15 cm, x = 25 cm, x = 35 cm, and x = 45 cm. This
reverse flow close to the bottom in run U1 was responsible,
at the beginning of the run, for the formation of the
upstream dune (shown in Figure 3) during the scouring
process described in section 3.1, because for series U, a
surface wave flow with a bottom recirculation developed.
When the velocity measurements were performed, that is, at
the end of the run, no sediment transport was observed for
series U, too, because the velocity values close to the
bottom were small.

4. Mathematical Simulation
4.1. Stability Analysis of a Surface Wave Jet
[18] The stability analysis is performed starting from the
motion equations considered in the streamline-centered s-n
system of reference [Rouse, 1938]. As a fundamental
hypothesis, we assume that the modulus of the mean
velocity vector does not depend on time, while the direction
of the velocity does. Such a hypothesis makes sense for
quasi-steady flow, in which the unsteadiness is due to
perturbations of an order of magnitude smaller than that
typical for the flow characteristics (discharge, downstream
water depth). The quantities required for the description of
the phenomenon are defined in Figure 7a. Moreover, we

Figure 7. Definition sketch for the mathematical model.

5 of 12

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

W02405

assume that the considered motion condition occurs in the


plane s-n. Such a hypothesis can be considered valid at
the start of the transition phenomenon, when the threedimensional effects are slight. Let V be the velocity vector,
defined as V = V(s,n)es, where V is the modulus, es is the
unit vector tangent to the instantaneous streamline, s is the
coordinate on it, and n is the coordinate along the direction
normal to the instantaneous streamline. The equations of
motion are
V

@V
1 @p
@z 1 @t

g 
@s
r @s
@s r @s


V2
@
p
z
g
@n
rg
R

where R is the radius of curvature of the instantaneous


streamline. The term  1r @t
@s accounts for head losses. As a
consequence, equation (1) gives the modified Bernoulli
equation
z

p V2 t

const
rg 2g rg

where
H x; t zb x h x; t

which is the same as that obtained by Montes and Chanson


[1998] but for the term 1r @t
@n, which accounts for the variation
of the tangential stress along the n coordinate.
[19] Now, adopting the geometrical hypothesis of Montes
and Chanson [1998], it is possible to transform the inde@
@
= cos(q)@z
, where q is the angle
pendent variable n as @n
between the horizontal axis and the local tangent to the
streamline, and to express all the geometrical quantities, R
and cos(q), as known functions of the respective bottom and
free surface quantities. As a consequence, the curvature of a
given streamline is expressed by means of the bottom and
free surface curvature.
[20] At last, introducing the dimensionless variable h,
defined as
h

z  zb x
hx; t

H0

@H 0
dzb
@2H
d 2 zb
; H 00 2 ; z00b 2
; zb
dx
dx
@x
@x

t tfs hk tb 1  hk

where k is a positive constant of order unity [Montes and


Chanson, 1998]. The coefficients appearing in equation (6)
account for the curvature of the streamlines and are defined
as
hH 00
hz00b
i ; a2 
a1  h
 2
1 z0b
1 H 0 2

10a

where tfs and tb are the tangential stresses on the free


surface and on the bottom, respectively. The former can be
neglected, while the latter can be expressed as tb = ru2* =
R1
U 2 n2
rg 1 m , where U = Vdh is the mean velocity and nm is the
h3

Mannings coefficient.
[22] In the second zone, the motion tends to become more
regular, and the stress law can be given as [Rodi, 2000;
Nezu and Nakagawa, 1993]
t


r 2m dV
dV
n
h h dh
dh

10b

where in this zone, dV


dh is always positive and m is the
mixing length, given by the following expression, considering a polynomial approximation of the Van Driest damping
function [Rodi, 2000; Nezu and Nakagawa, 1993]:
m khh

X
i

h
i
dV 1 dt
a1 hk a2 1  hk V 2 V

0
dh r dh

that is, such coefficients are defined in term of free surface


and bottom curvature.
[21] In order to obtain velocity profiles from equation (6),
it is necessary to describe the behavior of the stress along
the coordinate h. To this aim, it is useful to consider two
different zones of the scour region (Figure 6): zone 1 where
the flow is expanding (upstream of the abscissa of maximum scour depth), where the term which accounts for the
curvature of the streamlines in equation (6) is not negligible
with respect to the other two, and zone 2 where the flow is
accelerating (downstream of the abscissa of maximum scour
depth), where the term which accounts for the curvature of
the streamlines in equation (6) is less important than the
other two terms. In the first zone, the stress can be described
assuming that the same hypothesis is valid for the geometrical quantities, that is, as a combination of the stress on the
free surface and on the bottom,

equation (4) becomes

and

Inserting equation (3) in (2), the following equation for the


determination of V is obtained:
V2
@V 1 @t

V
@n r @n
R

W02405

"



#
1i hu* i
h
nA
i!

11

where k = 0.43 is the Von Karman constant; n is the


kinematic viscosity of the water; A is a damping factor,
which for open channel
q flows is equal to 26 [Nezu and Rodi,
1986]; and u* = nm g1 U is the friction velocity. Equation (6)
h3

will then be considered in the two different zones:


8h
i
i
dV 1 h
>
>
a1 hk a2 1  hk V 2 V
k tfs hk1 tb 1  hk1 0; zone 1
>
>
dh r
<
8
9
"

 #!2
= 1 dV
X
> dV
d <
1i hu* i
1 dV
>
>

h
n
0; zone 2
>
: V dh dh : khh
; h dh
i!
nA
h
dh
i

12
6 of 12

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

W02405

W02405

Figure 8. Comparison between the measured and predicted profiles of the u velocity component at four
different locations for test U1.

The general solution of equations (12) can be expressed as a


power series
V

ai hi

13

whose coefficients ai are defined by means of recurrence


relations obtained substituting solution (13) in the differential equations (12) and equating terms with the same
power of h. In particular, the first differential equation (12)
is a first-order equation; then only an integration constant
(corresponding to the first coefficient a0) has to be
determined. This latter will be determined by imposing
that the integral of (13), calculated with respect to h, is
equal to the depth averaged mean velocity U:

Z1
0

Vdh

N
X
i0

ai
i 1

[23] The analytical expressions of the horizontal and


vertical velocity components as well as the pressure are
given by
v V sin q

N
X

!
a i hi

i0

u V cosq

3
z0b

H
6
7
k
 41  hk q
 0 2 h q2 5 15
0
1 zb
1 H

N
X

!
ai h

i0

v
0
12
u
u
0
0
u
zb
Hb
B
C
k
k
u
t1  @1  h q
 0 2 h q
 0 2 A
1 zb
1 Hb

16

14

The second differential equation (12) is a second-order


equation; then two integration constants (corresponding to
the coefficients a0 and a1) have to be determined. These
latter are obtained imposing the following two conditions:
the integral of (13), calculated with respect to h, is equal to
the depth-averaged mean velocity U, and the derivative
of V,

u
d50
dV 
calculated on the bottom h = h , is equal to dh  d50 h kd*50
h

(i.e., it is assumed that the velocity profile has the same


inclination of the logarithmic profile).

Z1 ( "
z00 x; t
r 1  xk * b  2 +:
p rgh1  h  h
1 z0b
h
#
H 00 x; t
V 2 dx
xk
1 H 0 x; t 2

17

In Figures 8a 8d a comparison between the measured and


predicted velocity profiles of the horizontal velocity
component at four different locations, x = 5 cm (Figure 8a),
x = 65 cm (Figure 8b), x = 105 cm (Figure 8c), and x = 165 cm
(Figure 8d) for test U1 is shown. The agreement between

7 of 12

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

W02405

Table 2. Experimental Configurations: Downstream Froude


Number Frd, Measured Wavelength of the Free Surface Undulations
l0, Measured Amplitude of the Free Surface Undulations h0, and
Scour Wavelength ls
Run

Frd

l0 , m

h0, m

ls , m

Stability

T1
T2
T3
U1
U2
U3

0.15
0.17
0.18
0.14
0.15
0.16

0.12
0.15
0.18
0.35
0.25
0.35

0.0045
0.0030
0.0020
0.0030
0.0040
0.0035

2.2
2.0
2.8
2.2
2.3
2.9

Unstable
Unstable
Unstable
Stable
Stable
Stable

measured and predicted velocity profiles is good, showing


the validity of the assumed hypothesis.
[24] Expressions (15), (16), and (17) depend on zb(x),
h(x,t), U(x,t), and H(x,t) = zb(x) + h(x,t) as well as on the
Mannings coefficient. In this work zb(x) is obtained by
interpolating experimental data, while the Mannings coefficient is calculated as [Marchi and Rubatta, 1980]
1

nm

2d50 6
p
24 g

18

base solution h0(x), U0(x), while the second, known as the


base solution, permits us to determine the evolution of the
perturbations flow. The parameter g, which can be
considered as an eigenvalue of the linear and homogeneous
perturbation equations, gives information on the stability of
the base solution: If the real part of g is positive,
perturbations imposed on the base solution will tend to
destroy the base solution, and then this latter is considered
unstable. On the contrary, if the real part of g is negative,
perturbations imposed on the base solution will tend to
decrease with time, and the base solution is considered
stable. Of course, such conclusions are considered valid in
the framework of a linear stability analysis; that is, they are
valid in the early stage of the perturbations.
4.2. Computational Considerations and Analysis
[25] The differential equations for the determination of
the base flow and the perturbations are omitted for the sake
of simplicity. The aim of this work is to investigate the
stability of a given experimental condition, which is represented by the base flow, so this latter is assumed to be
known, interpolating experimental results. More precisely,
the free surface profile H0(x) = h0(x) + zb(x) and the bottom
profile z b (x) are measured along the x direction,
corresponding to the centerline of the experimental channel,
while the mean velocity profile U0(x) is obtained as

Finally, evolution equations for h(x,t),U(x,t) can be obtained


considering the integral forms of the continuity and
momentum equation [Montes and Chanson, 1998]
@h

@t

Z1

Vdh 0

19


Z1  2
@U
@
V
p
lU jU j
g
dh 
hh zb
@t
@x
rg
8<
2g

U0 x

Q
Q

bh0 x bH0 x  zb x

20

21

U x; t U0 x U1 xegt

22

It is supposed that the perturbations are small quantities


with respect to those of the base flow. Then, substituting
expressions (21) and (22) in equations (19) and (20), it is
possible to discard nonlinear terms with respect to the
perturbations. Imposing that the expressions (21) and (22)
satisfy equations (19) and (20), two distinct differential
systems are obtained. The first can be used to determine the



2px
l0


 

zd  jzmax j
zd jzmax j
2p x  x0

sin
zb x
ls
2
2

h x; t h0 x h1 xegt

23

where Q is the discharge, considered constant with respect


to time.
[26] Interpolating expressions for H0(x), zb(x) adopted in
this work are the following:
H0 x hd h0 sin

where l, and < are the friction factor and the hydraulic
radius, respectively, defined in the usual way for open
channel flow with a mobile bed [Marchi and Rubatta, 1980;
Garde and Ranga Raju, 1977]. Equations (19) and (20) are
the classic De Saint Venant equations for free surface
flows, in which curvature effects connected to the free
surface and bottom profiles have been accounted for.
Finally, the unknown functions h(x,t) and U(x,t) are defined
as the sum of a base flow and an unsteady perturbation with
a known dependency on time:

W02405

24

25

where l0 and h0 are the measured wavelength and amplitude


of the free surface undulations, respectively, defined as the
distance between the first two crests of the free surface and
half the distance between first crest and first throat over the
eroded bed, respectively. With reference to Figure 7b, zmax is
the maximum scour depth, zd is the dune height, x0 is the
abscissa where the scour depth is equal to zb(x0) = zd j2zmax j,
and ls is the scour hole wavelength defined in Figure 7b.
With such a definition of the bottom wavelength, the
expression (25) best fits the scour profile, as can be seen in
Figure 6. For the sake of simplicity, no decaying phenomena
were considered for the free surface undulations.
[27] Six experimental configurations were considered,
whose characteristics are shown in Table 2, where Frd is
the downstream Froude number, defined as
Frd

Q
p
bh0 gh0

26

In addition, in order to highlight the different characteristics


of the hydraulic jump stability, T1, T2, T3, U1, U2, and U3
experiments were performed with a flat bottom. In Table 3

8 of 12

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

W02405

Table 3. Experimental Configurations: Downstream Froude


Number Frd, Measured Wavelength of the Free Surface
Undulations l0, and Measured Amplitude of the Free Surface
Undulations h0
Run

Frd

l0 , m

h0, m

Stability

T1FB
T2FB
T3FB
U1FB
U2FB
U3FB

0.15
0.17
0.18
0.14
0.15
0.16

0.20
0.25
0.20
0.25
0.25
0.25

0.0029
0.0034
0.0031
0.0054
0.0144
0.0039

Stable
Stable
Stable
Stable
Stable
Stable

the experimental conditions performed with a flat bottom


are shown.
[28] The stability of the experimental configurations can
be investigated by solving the linear differential equations
for the perturbation flow h1(x), U1(x). Such differential
equations, though linear, have coefficients depending on
x, t, and an approximate procedure is adopted in order to
solve them. Such a procedure consists in assuming a Fourier
series development of the perturbation flow, in which the
fundamental wavelength of the perturbation flow coincides
with the wavelength of the free surface undulations, and in
substituting it in the perturbation equations. Imposing a
suitable solvability condition, an algebraic equation, whose
order depends on the number of modes adopted in the
Fourier expansion, is obtained for g. The real part of this
latter gives indications on the stability of the base flow.

W02405

[29] Results are shown in the stability diagrams presented


in Figures 9a 9f and 10a 10f, in which the horizontal axis
represents the wavelengths of the surface wave jet, and the
vertical axis represents the Froude number, defined with
the downstream characteristics of the flow. In Figures 9a 9f
the lines separate the unstable from the stable region; the
stability region is on the right of the lines. Figure 9a refers to
test U1FB, Figure 9b refers to test U2FB, and Figure 9c refers
to test U3FB; Figure 9d refers to test U1, Figure 9e refers to
test U2, and Figure 9f refers to test U3. The six stable
experimental configurations are represented by circles.
[30] In Figures 10a 10f the lines separate the unstable
from the stable region; the stability region is on the right of
the lines. Figure 10a refers to test T1FB, Figure 10b refers
to test T2FB, and Figure 10c refers to test T3FB; Figure 10d
refers to test T1, Figure 10e refers to test T2, and Figure 10f
refers to test T3. The six experimental configurations are
represented by circles.
[31] The horizontal lines limit the area over which it is
physically possible to have a hydraulic jump for the
experimental conditions considered. The stability diagram
is divided into three regions: the stable region on the right,
the unstable region on the left, and the no hydraulic jump
region at the bottom.
[32] The scoured bed influences deeply the surface wave
jet stability. Generally speaking, the presence of the scour
weakens the surface wave jet stability properties. As a matter
of fact, in Figures 10a 10f, relative to the comparison of
the stability diagrams with flat bottom (Figures 10a 10c)

Figure 9. Stability diagrams for tests (a) U1FB, (b) U2FB, (c) U3FB, (d) U1, (e) U2, and (f) U3.
9 of 12

W02405

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

W02405

Figure 10. Stability diagrams for tests (a) T1FB, (b) T2FB, (c) T3FB, (d) T1, (e) T2, and (f) T3.
and with scoured bed (Figures 10d 10f), the stability
characteristics of the oscillating jet are modified passing
from the flat bed to the scoured bed both because the
stability areas increase and because the characteristics of
the free surface undulations (amplitude and wavelength)
change, so the experimental point changes its position. In
particular, in the experimental configurations concerning
the transitional jet, the oscillating jet becomes stable if the
bed is flat.
[33] The instability of the surface wave jet can be
connected to the bed evolution. As a matter of fact, this
latter causes a nonequilibrium condition in the momentum
balance across the region occupied by the surface wave
jet;
* 2 in other
+ words, the value of the momentum flux
Q
h2
r bh gb 2 (where b is the flumes width) is not the
same in the extreme sections occupied by the surface wave
jet. Therefore the configuration of the flow adjusts itself to
seek a new equilibrium condition, causing the transition of
the surface wave jet to a submerged jet.
[34] Moreover, the presence of the scour increases the
dissipation: To this aim, it is interesting to compare the free
surface undulation amplitudes of tests U1 U3 and U1FB
U3FB in Tables 2 and 3 and Figure 11. The increased
dissipation is due to the fact that in the presence of a scour
hole, the flow expands itself. In conclusion, the stability
analysis of the modified De Saint Venant equations (19)
and (20) gave good results because, starting from the
experimental values of discharge, downstream water depth,
free surface undulation amplitude, length, and maximum

scour depth, it was able to highlight stability and instability


areas to which experimental points belong.

5. Conclusions
[35] Three series of laboratory experiments were carried
out to investigate local scour due to different kinds of water
jets, developing downstream of a drop followed by a rigid
apron. Three different kinds of jets, a submerged jet, a
surface wave jet, and an oscillating jet (defined as a
transition between submerged and surface wave jet), producing three different local scouring developments, were
observed. Measurements of the velocity fields show the
difference between the velocity field due to plunging flow
and to a surface wave flow. In the former, a submerged jet,
developing over the apron, produces a local scouring with a
downstream dune. In the latter, a surface wave jet with a
reverse flow close to the bottom, developing over the apron,
produces a local scouring with both an upstream and a
downstream dune. The scouring process due to the oscillating jet is a superimposition of the scouring process due to a
submerged jet and a surface wave jet.
[36] A mathematical model, derived starting from the
momentum equation, in the s, n coordinate system, permitted us to obtain analytical expressions for the horizontal
and vertical velocity components and for the pressure. To
this aim, the assumption of a suitable law for the stress
distribution along the vertical coordinate was important.
The stability analysis of the surface wave jet on both a flat
and a scoured bed, performed by a modified version of

10 of 12

W02405

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

hs
h0
H0
k
l0
ls
m
L
Lp
Ls
m
n
nm
p
Q
R
<
s
U
U0
U1
Figure 11. Free surface profiles for surface wave jets and
oscillating jet with both (a) rigid and (b) scoured bed.
De Saint Venant equations, permits us to classify the
experimental configuration considered. Interesting results
are connected with the link between theoretical and experimental analysis on the scouring processes due to a surface
wave flow. First, the mathematical model was able to
reproduce velocity profiles in different locations along
the scour, revealing the validity of the assumptions made.
Second, the comparison between the analysis of the
surface wave flow on flat and scoured beds highlighted
the influence of the scoured bed on the stability of the
surface wave flow. Generally speaking, the scoured bed
weakens the stability properties of the surface wave flow,
so that stable configurations with a flat bed are unstable
with a scoured bed. Moreover, the presence of the scour
increases the dissipation: As a matter of fact, larger surface
wave amplitudes are found with a flat bed than a scoured
bed. This fact is easily explained by the expansion of
the flow occurring with the scoured bed. The stability
diagrams obtained by the stability analysis confirmed all
the experimental observations.

u
u*
v
V
V
z
zb
zb0
zmax
ai
a0
g
h
h0
q
k
l
n
r
rs
t
tfs
tb

W02405

water depth over the drop.


downstream water depth over the rigid bottom.
water depth.
coefficient.
measured wavelength of the free surface undulations.
measured wavelength.
mixing length.
length of the sediment recess section.
apron length.
maximum scour length.
coefficient.
coordinate along the direction normal to the
instantaneous streamline at s = 0.
Mannings coefficient.
pressure.
water discharge.
instantaneous radius of curvature of the streamline.
hydraulic radius.
coordinate on the instantaneous streamline.
depth-averaged mean velocity along the x direction.
depth-averaged mean base velocity along the
x direction.
depth-averaged mean perturbation velocity along the
x direction.
mean velocity along the x direction.
friction velocity.
mean velocity along the z direction.
mean velocity vector.
mean velocity modulus.
coordinate along the vertical axis.
bottom depth.
scour depth in correspondence of the end of the
apron.
maximum scour depth.
coefficients.
coefficient.
eigenvalue of the linear and homogeneous perturbation equations.
dimensionless variable.
measured amplitude of the free surface undulations.
angle between the horizontal axis and the local
tangent to the streamline.
Von Karman constant.
friction factor.
kinematic viscosity of the water.
water density.
sediment density.
tangential stress.
tangential stress on the free surface.
tangential stress on the bottom.

Notation
A
a1
a2
b
d50
es
Dp
Frd
Fru
g
h

damping factor for Van Driests damping function.


coefficient.
coefficient.
flumes width.
mean sediment size.
unit vector tangent to the instantaneous streamline.
height of drop.
downstream Froude number.
upstream Froude number.
acceleration of gravity.
instantaneous water depth.

[37] Acknowledgment. Laboratory activities were supported by Italian funding COFIN 2002 2004, Modellazione di processi idrodinamici in
sistemi fluidi stratificati, di correnti a superficie libera e in pressione a
celerita` variabile, section Modellazione sperimentale di fenomeni erosivi
localizzati nelle correnti a superficie libera.

References
Adduce, C. (2004), Local scour downstream of a turbulent jet, Ph.D. thesis,
167 pp., Univ. of Rome Roma Tre, Rome, Italy, 20 Oct.
Adduce, C., and P. Mele (2004), Local scour by submerged turbulent
jets, paper presented at International Conference on Hydroscience and
Engineering 2004, Int. Assoc. for Hydraul. Res. and Eng., Brisbane,
Queensland, Australia, 31 May to 3 June.

11 of 12

W02405

ADDUCE AND LA ROCCA: LOCAL SCOURING BY TURBULENT JETS

Adduce, C., M. La Rocca, and P. Mele (2004), Local scour downstream of


grade control structures, paper presented at Hydraulics of Dams and
River Structures 2004, Int. Assoc. for Hydraul. Res. and Eng., Tehran,
26 28 April.
Alfonsi, G. (2001), Analysis of streamwise velocity fluctuations in turbulent pipe flow with the use of an ultrasonic Doppler flowmeter, Flow
Turbul. Combust., 67, 137 142.
Bormann, N. E., and P. Y. Julien (1991), Scour downstream of gradecontrol structures, J. Hydraul Eng., 117, 579 594.
Breusers, H. N. C., and A. J. Raudkivi (1991), Scouring, A. A. Balkema,
Brookfield, Vt.
Bukreev, V. I. (2001), Undular jump in open-channel flow over a sill,
J. Appl. Mech. Tech. Phys., 42(4), 596 602.
Chatterjee, S. S., S. N. Ghosh, and M. Chatterjee (1994), Local scour due to
submerged horizontal jet, J. Hydraul. Eng., 120, 973 992.
Dey, S., and B. Westrich (2003), Hydraulics of submerged jet subject to
change in cohesive bed geometry, J. Hydraul. Eng., 129, 44 53.
Eckert, S., and G. Gerbeth (2002), Velocity measurements in liquid sodium
by means of ultrasound Doppler velocimetry, Exp. Fluids, 32, 542 546.
Farhoudi, J., and K. V. H. Smith (1985), Local scour profiles downstream of
hydraulic jump, J. Hydraul. Res., 23, 343 358.
Fritz, H. M., and W. H. Hager (1998), Hydraulics of embankment weirs,
J. Hydraul. Eng., 124, 963 971.
Garde, R. J., and K. G. Ranga Raju (1977), Mechanics of Sediment Transport and Alluvial Stream Problems, John Wiley, Hoboken, N. J.
Gijs, J., and J. C. M. Hoffmans (1998), Jet scour in equilibrium phase,
J. Hydraul. Eng., 124, 430 437.
Hager, W. H. (1992), Energy Dissipators and Hydraulic Jumps, Springer,
New York.
Hassan, N. M. K. N., and R. Narayanan (1985), Local scour downstream of
an apron, J. Hydraul. Eng., 111, 1371 1385.
Hoffmans, G. J. C. M. (1997), Jet scour in equilibrium phase, J. Hydraul.
Eng., 124, 430 437.
Hoffmans, G. J. C. M., and H. J. Verheij (1997), Scour Manual, A. A.
Balkema, Brookfield, Vt.
Hurther, D. (2001), 3-D acoustic Doppler velocimetry and turbulence in
open-channel flow, Ph.D. thesis, Ecole Polytech. Fed. de Lausanne,
Lausanne, Switzerland.
Ito, T., Y. Tsuji, H. Nakamura, and Y. Kukita (2001), Application of ultrasonic velocity profile meter to vortex shedding and empirical eigenfunctional analysis, Exp. Fluids, 31, 324 335.
Kawagoshi, N., and W. H. Hager (1990), B-jump in sloping channel II,
J. Hydraul. Res., 28, 461 480.
Kurniawan, A., M. S. Altinakar, and W. H. Graf (2001), Flow pattern of an
eroding jet, paper presented at 29th IAHR Congress, Int. Assoc. for
Hydraul. Res., Beijing, 16 21 Sept.
Longo, S., I. J. Losada, M. Petti, N. Pasotti, and J. L. Lara (2001), Measurements of breaking waves and bores through a USD velocity profiler,
Tech. Rep. UPR/UCa_01_2001, Univ. of Parma, Parma, Italy.

W02405

Marchi, E., and A. Rubatta (1980), Meccanica dei Fluidi: Principi e


Applicazioni Idrauliche, Unione Tiopgrafico-Editrice Torinese, Bologna,
Italy.
Mazurek, K. A., N. Rajaratnam, and D. C. Sego (2003), Scour of cohesive
soil by submerged plane turbulent wall jets, J. Hydraul. Res., 41, 195
206.
Montes, J. S., and H. Chanson (1998), Characteristics of undular hydraulic
jumps: Experiments and analysis, J. Hydraul. Eng., 124, 192 205.
Nezu, I., and H. Nakagawa (1993), Turbulence in Open-Channel Flow,
IAHR-AIRH Monogr. Ser., A. A. Balkema, Brookfield, Vt.
Nezu, I., and W. Rodi (1986), Open-channel flow measurements with a
laser Doppler anemometer, J. Hydraul. Eng., 112, 335 355.
Ohtsu, I., and Y. Yasuda (1991a), Hydraulic jump in sloping channels,
J. Hydraul. Eng., 117, 905 921.
Ohtsu, I., and Y. Yasuda (1991b), Transition from supercritical to subcritical
flow at an abrupt drop, J. Hydraul. Res., 29, 309 328.
Rajaratnam, N. (1965), The hydraulic jump as wall jet, J. Hydraul. Div. Am.
Soc. Civ. Eng., 91, 107 131.
Rajaratnam, N. (1976), Turbulent Jet, Elsevier, New York.
Rajaratnam, N. (1981), Erosion by plane turbulent jets, J. Hydraul. Res., 19,
339 358.
Rajaratnam, N., and V. Murahari (1974), Flow characteristics of sloping
channel jump, J. Hydraul. Div. Am. Soc. Civ. Eng., 100, 731 740.
Rodi, W. (2000), Turbulence Models and Their Application in Hydraulics,
IAHR-AIRH Monogr. Ser., A. A. Balkema, Brookfield, Vt.
Rolland, T. (1994), Development dune instrumentation Doppler ultrasonore adaptee a` letude hydraulique de la turbulence dans les canaux,
Ph.D. thesis, Ecole Polytech. Fed. de Lausanne, Lausanne, Switzerland.
Rouse, H. (1938), Fluid Mechanics for Hydraulic Engineering, McGrawHill, New York.
berfallstrahlen, WasserSchoklitsch, A. (1932), Kolkbildung unter U
wirtschaft, 24, 341 343.
Takeda, Y. (1995), Velocity profile measurements by ultrasonic Doppler
method, Exp. Therm. Fluid Sci., 10(4), 444 453.
Takeda, Y. (1999), Ultrasonic Doppler method for velocity profile measurements in fluid dynamics, Exp. Fluids, 26, 177 178.
Willemetz, J. C. (1990), Etude quantitative de lhemodynamique de vaisseaux profonds par echographie Doppler ultrasonore, Ph.D. thesis, Ecole
Polytech. Fed. de Lausanne, Lausanne, Switzerland.
Willemetz, J. C. (2001), DOP 2000 users manual, Signal Process. SA,
Lausanne, Switzerland.
Wu, S., and N. Rajaratnam (1998), Impinging jet and surface flow regimes
at drop, J. Hydraul. Res., 36, 69 74.




C. Adduce and M. La Rocca, Dipartimento di Scienze dellIngegneria


Civile, Universita` degli Studi Roma Tre, Via Vito Volterra 62, I-00146,
Roma, Italy. (adduce@uniroma3.it)

12 of 12

You might also like