Vous êtes sur la page 1sur 12

Color profile: Disabled

Composite Default screen

33

Interactions between cavity flow and main stream


skimming flows: an experimental study
C.A. Gonzalez and H. Chanson

Abstract: In the last two decades, research on the hydraulics of skimming flows down stepped chutes was driven by
needs for better design guidelines. Skimming flows are characterized by significant momentum transfer from the main
stream to the recirculation zones. Investigations are difficult because of the complex nature of the flow, the strong flow
aeration, and the interactions between entrained air and turbulence. This study provides a comprehensive database on
main stream and cavity flow interactions in skimming flows down a stepped chute. Measurements were conducted on a
large facility (α = 15.9°, h = 0.05 and 0.1 m, W = 1 m) with precise instrumentation based upon a Froude similitude.
Air–water velocity and turbulence measurements demonstrated a well-defined mixing layer developing downstream of
each step edge in which the velocity profiles had the same shape as classical monophase flow results. A comparative
analysis of air–water flow properties for Froude similar flow conditions showed a good agreement between the two
step heights in terms of dimensionless distributions of air content, velocity, and turbulence intensity, but dimensionless
bubble count rates, turbulence levels, and bubble chord sizes were improperly scaled.
Key words: stepped chute, skimming flow, flow recirculation, momentum exchange, physical modelling.
Résumé : Durant les dernières deux décennies, les recherches sur l’hydraulique des écoulements extrêmement turbu-
lents sur coursiers en marches d’escalier ont été dictées par les besoins de l’industrie pour de meilleures lignes directri-
ces de conception. Les écoulements extrêmement turbulents (« skimming flows ») sont caractérisés par un transfert
important de quantité de mouvement entre l’écoulement principal et les zones de recirculation dans les marches.
L’étude en profondeur de ces processus est compliquée par la complexité de l’écoulement, la présence importante de
bulles et les interactions dynamiques entre les bulles d’air et les structures tourbillonnaires. Dans cette étude, on pré-
sente une base de données expérimentales complète sur les interactions entre l’écoulement principal et les zones de re-
circulation, pour des écoulements extrêmement turbulents sur un coursier en marches d’escalier. Ce travail a été réalisé
dans un canal de grande taille (α = 15,9°, h = 0,05 et 0,1 m, W = 1 m), avec une instrumentation très précise, en se
basant sur la loi de Reech-Froude. Les mesures de vitesses du mélange air–eau et d’intensité de turbulence démontrent
l’existence d’une couche de mélange bien définie en aval de chaque arête de marche (« step edge »); les profils de vi-
tesse présentent une analogie avec les couches de mélange en écoulements monophasiques. En se basant sur la loi de
Reech-Froude, une étude avec deux hauteurs de marches démontre des résultats adimensionels comparables, en termes
de taux de vide, vitesse et intensité de la turbulence, mais les taux de bulles d’air adimensionels, les niveaux de turbu-
lence et les tailles de bulles d’air n’avaient pas le bon ordre de grandeur.

Mots clés : coursier en marches d’escalier, écoulement extrêmement turbulent, recirculation, échange de quantité de
mouvement, modélisation physique.
[Traduit par la Rédaction] Gonzalez and Chanson 44

Introduction bottom formed by step edges (Rajaratnam 1990) (Fig. 1).


Beneath the three-dimensional cavity vortices develop and
During the last two decades, research on the hydraulics of recirculation is maintained through the transmission of shear
stepped chutes was driven by needs for better design guide- stress from the main stream (Fig. 2). Small-scale vorticity is
lines (Chanson 1995; Ohtsu and Yasuda 1998; Minor and also generated at the corner of the steps. Skimming flows
Hager 2000; Chanson 2001). Most research was conducted are characterized by very significant form losses and mo-
for skimming flows corresponding to the largest discharges mentum transfer from the main stream to the recirculation
per unit width. That is, the waters flow down a stepped zones. There is an obvious analogy with skimming flows
channel as a coherent stream skimming over the pseudo- past large roughness elements and cavities (Townes and

Received 2 May 2003. Revision accepted 14 July 2003. Published on the NRC Research Press Web site at http://cjce.nrc.ca on
13 January 2004.
C.A. Gonzalez and H. Chanson.1 Department of Civil Engineering, The University of Queensland, Brisbane, QLD 4072, Australia.
Written discussion of this article is welcomed and will be received by the Editor until 30 June 2004.
1
Corresponding author (e-mail: h.chanson@uq.edu.au).

Can. J. Civ. Eng. 31: 33–44 (2004) doi: 10.1139/L03-066 © 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:06 PM
Color profile: Disabled
Composite Default screen

34 Can. J. Civ. Eng. Vol. 31, 2004

Fig. 1. Skimming flow downs Camp Dyer Diversion Dam spillway — unprotected RCC stepped spillway over an old masonry weir
(courtesy of the US Bureau of Reclamation).

Fig. 2. Definition sketch of a skimming flow.

Sabersky 1966; Knight and Macdonald 1979; Djenidi et al. conducted on a large facility (α = 15.9°, h = 0.05 and 0.1 m,
1994; Elavarasan et al. 1995; Tantirige et al. 1994; Manso W = 1 m, where α is the channel slope, h is the height of
and Schleiss 2002). In stepped chutes, however, little re- steps measured vertically, and W is the channel width) with
search was conducted on the interactions between the main precise instrumentation. The results provide a better under-
stream and the cavity recirculation, with the exception of standing of the momentum exchange processes.
preliminary experiments by Boes (2000), Chanson and
Toombes (2002a), and Matos et al. (2001), and some crude
modelling by Chanson et al. (2000, 2002). Investigations are Similitude and dimensional analysis
difficult because of the complex nature of the flow, the In skimming flows down a stepped chute, flow resistance
strong flow aeration, and the interactions between entrained is primarily step form drag. Free-surface aeration is very in-
air and turbulence. tense, and its effects cannot be neglected. Analytical and nu-
It is the purpose of this study to provide a comprehensive merical studies of skimming flows are difficult because of
database on main stream and cavity flow interactions in the number and complexity of the relevant equations. Exper-
skimming flows down a stepped chute. Measurements were imental investigations are preferred, and this study is no ex-

© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:07 PM
Color profile: Disabled
Composite Default screen

Gonzalez and Chanson 35

ception. In a channel made of flat horizontal steps, a with high-shutter-speed digital equipment, i.e., a digital video-
complete dimensional analysis yields camera handycam Sony™ DV-CCD DCR-TRV900 (speed:
25 frames/s, shutter: 1/4 to 1/10 000 s) and a digital camera
V u′ v′ d
[1] C, , , , ab , Olympus™ Camedia C-700 (shutter: 1/2 to 1/1 000 s).
gde V V de Air–water flow properties were measured using a double-
tip conductivity probe (Ø = 0.025 mm for each sensor). The
 
q gµ 4w q w W k s′  probe sensors were aligned in the flow direction and excited
= F  l ; ; ρw w ;
x y
; ; ; α; by an air bubble detector (AS25240). The probe signal was
 de de µ w ρwσ 3 gh 3 h h 
 scanned at 20 kHz per sensor for 20 s. Most measurements
were conducted with a probe tip separation of ∆x = 8 mm in
where C is the void fraction, V is the velocity (metres per the streamwise direction. (The exact distance ∆x was mea-
second), de is an equivalent clear-water depth (metres), g is sured with a microscope Beck/London Model 2294 with an
the acceleration due to gravity (metres per second squared), error of less than 0.00217 mm.) The shear flow region im-
u′ is the root mean square of the axial component of turbu- mediately downstream of the outer step edge was character-
lent velocity (metres per second), v′ is the root mean square ized by intense turbulent shear and recirculation. A few
of the lateral component of turbulent velocity (metres per measurements in that flow region (xs/Lcav < 0.5 and y < 0)
second), dab is a characteristic bubble size (metres), xl is the were performed with a probe sensor spacing of ∆x =
longitudinal distance measured in the flow direction 3.18 mm, where xs is the streamwise distance from the step
(metres), y is the distance measured normal to the pseudo- edge and Lcav is the step cavity length (metres). (Fig. 2). The
bottom formed by the step edges (metres), qw is the water shorter probe tip spacing authorized better cross-correlations
discharge per unit width (square metres per second), µw and among probe tip signals.
ρw are the dynamic viscosity (pascal second) and density of The translation of the probes in the direction normal to
water (kilograms per cubic metre), respectively, σ is the sur- the channel invert was controlled by a fine adjustment travel-
face tension (newtons per metre), α is the angle between the ling mechanism connected to a Mitutoyo™ digimatic scale
horizontal and the pseudo-bottom formed by the step edges, unit. The error on the vertical position of the probe was less
and ks′ the equivalent sand roughness height of the step faces than 0.1 mm.
(Fig. 2). For air–water flows, de is usually defined as
y =Y90 Experimental flow conditions
[2] de = ∫ (1 − C ) dy Experimental investigations were conducted for dimen-
sionless flow rates dc /h ranging from 0.6 to 3.2, where dc is
y=0
the critical depth. For dc /h < 0.6, a succession of free-falling
where Y90 is the characteristic distance for C = 0.9. nappes was observed. For 0.6 < dc /h < 1.25, the flow exhib-
For geometrically similar models, it is impossible to sat- ited a chaotic flow behaviour associated with strong droplet
isfy simultaneously more than one similitude, and scale ef- ejection processes downstream of the inception point of free-
fects will exist when one or more π-terms have different surface aeration, i.e., transition flow regime. For dc /h ≥ 1.3,
values in the model than the prototype. For example, in the flow skimmed over the pseudo-bottom formed by the
small-sized models based upon a Froude similitude, the air step edges, i.e., skimming flow regime. Irregular-cavity fluid
entrainment process may be affected by significant scale ef- ejections were observed that were evidences of momentum
fects (Wood 1991; Chanson 1997). Similarly, for stepped transfer between the main stream and the cavity flows. The
chute studies based upon a Froude similitude, scale effects recirculating fluid flowed outwards into the main stream and
in terms of flow resistance are small when the Reynolds was replaced by new fluid. The ejection and inflow pro-
number and step height satisfy, i.e., ρw qw/µw > 2.5 × 104 and cesses took place predominantly near the downstream end of
h > 0.02 m (Chanson et al. 2002). In the present study, a the cavity.
Froude similitude was used as for most open channel flow The present study focused on the highly aerated skimming
studies (Henderson 1966; Chanson 1999). Detailed air–water flow regime. For five flow rates with h = 0.1 m and seven
measurements were conducted in a large-sized facility to en- flow rates with h = 0.05 m, detailed air–water flow measure-
sure that the experimental results might be up-scaled with ments were conducted at all outer step edges downstream of
negligible scale effects (Table 1). inception and at several positions xs/Lcav in and above the
recirculation cavities.
Experimental apparatus and instrumentation
Air–water flow properties at step edges
Experiments were conducted at The University of Queens-
land in a 1-m wide channel previously used by Chanson and At the upstream end of the cascade, the flow was smooth
Toombes (2002a). The new test section was 4.2 m long and and no air entrainment occurred. After a few steps the flow
consisted of a broad crest followed by 9 identical steps of was characterized by a strong air entrainment. A similar lon-
0.10 m height or 18 steps of 0.05 m height (Table 1). The gitudinal pattern is seen in Fig. 1. Downstream, the two-
chute slope was α = 15.9° (l = 0.35 and 0.175 m, respec- phase flow behaved as a homogeneous mixture. The exact
tively). The flow rate was supplied by a pump controlled location of the interface became undetermined. There were
with an adjustable frequency AC motor drive. The discharge continuous exchanges of air–water and momentum between
was measured from the upstream head above crest with an the main stream and the atmosphere. The air–water flow
accuracy of about 2%. Flow visualizations were conducted consisted of a bubbly flow region (C < 30%), a spray region

© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:07 PM
Color profile: Disabled
Composite Default screen

36 Can. J. Civ. Eng. Vol. 31, 2004

Table 1. Detailed experimental investigations of air entrainment in stepped chutes.


Reference α (°) qw (m2/s) h(m) Flow regime Instrumentation Remarks
Chanson and 3.4 0.038–0.163 0.143 Nappe flow Single-tip conductivity L = 24 m; W = 0.5 m; super-
Toombes probe (Ø = 0.35 mm) critical inflow (0.03-m noz-
(1997, 2002c) zle thickness)
Tozzi et al. 52.2 0.23 0.053 Skimming flow Conductivity probe Inflow: uncontrolled smooth
(1998) WES ogee crest followed
by smaller first steps
Chamani and 51.3 and 0.07–0.2 0.313–0.125 Skimming flow Conductivity probe and W = 0.30 m; inflow: uncon-
Rajaratnam 59 flushed Pitot tube (Ø trolled smooth WES ogee
(1999) = 3.2 mm) crest
Matos (2000) 53.1 0.08–0.2 0.08 Skimming flow Conductivity probe and W = 1 m; inflow: uncontrolled
flushed Pitot tube (Ø WES ogee crest, with small
= 3.2 mm) first steps built in the ogee
development
Toombes and 3.4 0.08–0.136 0.143 Nappe flow Double-tip conductivity L = 3.2 m; W = 0.25 m;
Chanson probe (Ø = 0.025 mm) supercritical inflow (nozzle
(2000) thickness 0.028 to 0.040
m); ventilated steps
Boes (2000) 30 and 0.047–0.38 0.023–0.09 Skimming flow Double-tip optical fibre W = 0.5 m; inflow: pressur-
50 probe RBI (Ø = 0.1 ised intake
mm, 2.1 mm spacing
between sensors)
Ohtsu et al. 55 0.016–0.03 0.025 Skimming flow Single-tip optical fibre W = 0.3 m; inflow: uncon-
(2000) probe trolled broad crest
Chanson and 21.8 0.06–0.18 0.1 Transition and Double-tip conductivity L = 3.0 m; W = 1 m; inflow:
Toombes skimming probe (Ø = 0.025 mm) uncontrolled broad crest;
(2002a, flows experiments TC200
2002b)
15.9 0.07–0.19 0.1 Transition and Double-tip conductivity L = 4.2 m; W = 1 m; inflow:
skimming probe (Ø = 0.025 mm) uncontrolled broad crest;
flows experiments TC201
Present study 15.9 0.020–0.200 0.05 Transition and Double-tip conductivity L = 4.2 m; W = 1 m; inflow:
skimming probe (Ø = 0.025 mm) uncontrolled broad crest;
flows experiments CG202
0.075–0.220 0.10 Transition and Including detailed measure-
skimming ments between step edges
flows
Note: L, chute length; W, chute width.

(C > 70%), and an intermediate flow structure for 0.3 < C < 50%, while the relationship between bubble count rate and
0.7 (Fig. 2). Waves and wavelets could propagate along the void fraction was quasi-parabolic (Fig. 4b).
free surface (Toombes 2002). Two phase flow velocity distributions are presented in
At the step edges, the advective diffusion of air bubbles may Fig. 3b in terms of the time-averaged air–water velocity V
be described by an analytical model of air bubble diffusion and turbulence intensity Tu = u′/V. The latter was deduced
from the width of the cross-correlation function. The pro-
cessing technique was detailed in Chanson and Toombes
 3
y / Y90 (y / Y90 − 1/ 3)  (2002b). In skimming flows, the velocity data at step edges
[3] C = 1 − tan h 2  K′ − + compared favourably with a power law
 2Do 3Do 
  1/ N
V  y 
[4] = 
where K′ is an integration constant and Do is a dimensionless V90  Y90 
diffusivity term, which is a function of the mean void frac-
tion (Chanson and Toombes 2002a). The data were com- where V90 is the characteristic velocity for C = 90%
pared successfully with eq. [3] (Fig. 3a). Figure 3a also (Fig. 3b). Overall, N was found to be typically between 5
presents dimensionless distributions of bubble count rates and 12, but the data exhibited some longitudinal oscillations
Fdc/Vc, where F is the bubble count rate (hertz) defined as with a wave length of about 2 to 3 step cavity lengths. Such
the number of bubbles detected by the probe sensor per sec- longitudinal oscillations were also observed in terms of
ond and Vc is the critical flow velocity. For all flow rates, the mean air content Cmean and mean flow velocity Uw (Table 2)
data showed maximum bubble count rates for C ≈ 40% to where
© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:07 PM
Color profile: Disabled
Composite Default screen

Gonzalez and Chanson 37

Fig. 3. Air–water flow properties in skimming flow at step edges for qw = 0.0643 m2/s, dc/h = 1.5, h = 0.05 m: (a) void fraction and
dimensionless bubble count rate distributions — comparison with eq. [3], (b) air–water velocity distributions — comparison with
eq. [4], and (c) turbulence intensity distributions.

y =Y90

∫ (1 − C )Vdy Although Table 2 presents depth-averaged results for one


y=0
[5] Uw = y =Y90
flow rate only, the data were relatively typical of all results.
Some data of Boes (2000) and Matos (2000) showed similar
∫ (1 − C ) dy longitudinal oscillations in terms of depth-averaged air con-
y=0 tents. It is believed that these longitudinal waves were the
© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:07 PM
Color profile: Disabled
Composite Default screen

38 Can. J. Civ. Eng. Vol. 31, 2004

Fig. 4. Air–water flow properties in skimming flow between step edges for qw = 0.219 m2/s, dc/h = 1.7, h = 0.10 m, between step
edges 8 and 9, xs/Lcav = 0.4: (a) void fraction and air–water velocity distributions: comparison with eqs. [3] and [4] at step edge 8,
(b) dimensionless bubble count rate distributions: comparison with a parabolic law, and (c) turbulence intensity distributions.

© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:07 PM
Color profile: Disabled
Composite Default screen

Gonzalez and Chanson 39

Table 2. Longitudinal air–water flow properties for qw = Prandtl’s mixing length (Rajaratnam 1976). Goertler’s solu-
0.0643 m2/s, dc/h = 1.5, h = 0.05 m. tion of the equations of motion is
Step edge Cmean Uw/Vc N Remarks
V 1  y - y50  
8 — — — Inception point [7] = 1 + erf  K 
9 0.387 2.50 9.7 Vo 2   x s  
10 0.285 2.59 5.7
11 0.419 2.45 16 where y50 is the location where V/Vo = 0.5, K is a constant
12 0.274 2.58 4.6 inversely proportional to the rate of expansion of the mixing
13 0.421 2.48 9.8 layer, and erf is the error function given by
14 0.325 2.64 8.4
15 0.407 2.83 9.3 u
1
∫ e − t dt
2
16 0.346 2.60 8.0 [8] erf(u) =
17 0.368 2.62 7.9 π0
Note: Uw = qw/de; Vc, critical flow velocity.
where u and t are dimensionless variables used to define erf.
In monophase flows, a was found to be 0.084 and 0.09 for
result of strong interactions between vortex shedding down- the data sets of Liepmann and Laufer (1947) and Albertson
stream of each step edge and the free surface. et al. (1950), respectively, while K was equal to 11 for the
The distributions of turbulence intensity Tu showed high data of Liepmann and Laufer. For the experimental data pre-
turbulence levels across the entire air–water flow mixture, sented in Figs. 4 and 5, values of the coefficients a and K are
i.e., 0 ≤ y ≤ Y90 (Fig. 3c). The trend differed significantly summarized in Table 3. (These data were obtained from the
from turbulence intensity profiles observed in turbulent best data fit.) Along one-step cavity, the coefficient K in-
boundary layer flows (Schlichting 1979), although they are creased with xs towards monophase flow values (K = 11),
close to the earlier results of Chanson and Toombes (2002a) while the values of a decreased with xs towards the reported
on a 22° slope stepped chute. values for monophase flow (a ≈ 0.09).
Figure 5b demonstrates self-similarity of the velocity pro-
Air–water flow properties between step files. In this figure, the velocity data are presented as V/Vo
edges versus K(y – y50)/xs where Vo was selected such as Vo =
0.9V90. Experimental observations agreed well with both
Between step edges (0 < xs/Lcav < 1), air–water flow prop-
Tollmien’s and Goertler’s solutions.
erties exhibited significant differences, particularly for
The upper edge of the shear layer (located where V/Vo = 1)
y/Y90 < 0.3. Figures 4 and 5 illustrate a data set for one cav-
and the free velocity V = Vo/2 (y = y50) were recorded for all
ity flow. Between step edges, void fraction distributions
the locations in the cavity. The results are reported in Fig. 6.
showed greater flow aeration than at step edges (Fig. 4).
The experimental observations highlight the shape of the de-
Matos et al. (2001) reported a similar finding. Mean air con-
veloping shear layer downstream of each step edge. The
centrations calculated from the pseudo-bottom formed by
above finding provides means to develop a relationship for
the step edges (y = 0) were typically 20% to 30% larger than
the growth of the mixing layer and predict the mean velocity
those observed at the upstream and downstream step edges
distribution based upon a suitable shear-stress model.
(Table 3). It was proposed that air bubbles were trapped in
the large-scale vortical structures of the recirculation zone
by inertial effect (Matos et al. 2001). Turbulence levels
Dimensionless velocity distributions are shown in Fig. 5, Turbulence intensity distributions in the shear layers are
while characteristic parameters are summarized in Table 3 presented in Fig. 5c. The data showed very high levels of
for the same data set. Figure 5a shows longitudinal varia- turbulence in the shear flow. Maximum turbulent intensities
tions of the velocity distribution above the recirculation of more than 60% were observed. These values were consis-
zone. The data suggest a developing shear layer downstream tent with turbulence intensity measurements in plunging jet
of the singularity formed by the step edge. In Fig. 5b, the flows by Chanson and Brattberg (1998) and in wake flows
experimental data are compared with the theoretical solu- between rocks by Sumer et al. (2001). However, the present
tions of Tollmien and Goertler for plane turbulent shear lay- data were significantly larger than turbulence levels ob-
ers (Rajaratnam 1976; Schlichting 1979). For a free shear served in monophase developing shear flows. In monophase
layer, Tollmien’s solution of the equations of motion yields mixing layers, experimental data indicated maximum turbu-
lence levels (Tu)max = 15% to 20% for xs/dj ≤ 4 where dj is
V d   3  the jet flow thickness (Davies 1966; Sunyach and Mathieu
[6] =  − 0.0176eØ + 01337
. eØ/ 2 cos  Ø
Vo dØ   2  1969; Wygnanski and Fiedler 1970).
The present data suggested further greater turbulence inten-
 3  sities next to the downstream end of the cavity (xs/Lcav ≥ 0.5).
. eØ/ 2 sin 
+ 0687 Ø  For example, the maximum turbulence levels (Tu)max were
 2   about 0.8, 1, 1.1, and 1.5 for xs = 0.2, 0.3, 0.6, and 0.75, re-
spectively (Fig. 5c). The findings might be consistent with vi-
where Ø ∝ y/(axs); Vo is the free-stream velocity; and a is an sual observations of cavity fluid ejection and replenishment
empirical constant that equals (2lm2/xs2)1/3, where lm is taking place primarily next to the cavity downstream end.
© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:08 PM
Color profile: Disabled
Composite Default screen

40 Can. J. Civ. Eng. Vol. 31, 2004

Fig. 5. Air–water velocity distributions between step edges for qw = 0.219 m2/s, dc/h = 1.7, h = 0.10 m, between step edges 8 and 9:
(a) velocity distributions, (b) self-similarity: comparison between experimental data and the theoretical solutions of Tollmien and
Goertler, and (c) turbulence intensity distributions in the shear layer.

Discussion 0.10 m. A detailed comparison of the results obtained with


the two step heights for a similar flow rate showed good
Air–water flow similarity agreement between the two configurations in terms of
For two dimensionless flow rates (dc/h = 1.5 and 1.7), dimensionless distributions of air content, velocity, and tur-
identical experiments were repeated with h = 0.05 and bulence intensity, as well as in terms of mean air content

© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:09 PM
Color profile: Disabled
Composite Default screen

Gonzalez and Chanson 41

Table 3. Air–water flow properties between two adjacent step edges for qw = 0.219 m2/s, dc/h = 1.7, h = 0.10 m,
between step edges 8 and 9.
Goertler’s Tollmien’s
Step edge xs/Lcav Cmean Uw/Vc V90/Vc solution, K solution, a (Fmaxdc)/Vc
8 0 0.37 2.42 2.72 na na 30.4
0.2 0.33 2.27 2.75 2.46 1.192 26.2
0.25 0.35 2.31 2.66 2.32 0.975 29.3
0.3 0.39 2.29 2.64 2.36 0.809 26.9
0.4 0.42 2.33 2.69 2.73 0.615 24.5
0.5 0.46 2.37 2.67 5.16 0.487 26.9
0.6 0.53 2.34 2.63 4.17 0.417 24.0
0.75 0.51 2.34 2.69 5.94 0.325 25.1
0.8 0.47 2.31 2.75 4.02 0.297 26.7
9 1.0 0.39 2.37 2.75 na na 31.6
Note: Cmean, integrated between y = 0 and Y90; Lcav = 0.364 m; na, not applicable.

Fig. 6. Sketch of the developing shear layer and experimental data points for qw = 0.219 m2/s, dc/h = 1.7, h = 0.10 m, between step
edges 8 and 9.

Cmean, dimensionless flow velocity Uw /Vc, and air–water


flow velocity V90 /Vc. However, significant differences were
Turbulent shear stress
observed in terms of dimensionless bubble count rates
At each step, the cavity flow is driven by the developing
Fdc /Vc, while lesser maximum turbulence levels were noted
shear layer and the transfer of momentum across it (Figs. 2
with h = 0.05 m. At each step edge, measurements for h =
and 5). The equivalent boundary shear stress of the cavity
0.05 m showed lesser dimensionless bubble count rates by
flow equals the maximum shear stress τmax in the shear layer
about 30% to 50% than those observed for h = 0.10 m at an
that may be modelled by a mixing length model
identical step edge for the same dimensionless flow rate.
The finding suggests some scale effects in terms of bubble  ∂V 
counts and bubble sizes. [9] τmax = ρνt  
A comparative analysis of bubble chord size distributions  ∂y  y = y
50
for identical flow rate, location, and local void fraction
showed consistent differences between the two step heights where νt is the momentum exchange coefficient (Chanson et
that were not scaled at 2:1, implying that eq. [1] could not al. 2000). For Goertler’s solution of the equations of motion,
be approximated properly by a Froude similitude (e.g., the dimensionless pseudo-boundary shear stress equals
Fig. 7). Figure 7 compares bubble chord sizes recorded at
8τmax 2
the same dimensionless distance from the inception point [10] =
and for the same dimensionless flow rate with the two step ρ Vo2 πK
heights (h = 0.1 and 0.05 m). Figure 7a shows air chord size
distributions in 0.5 mm intervals for C ≈ 0.1. The data show where 1/K is the rate of expansion of the mixing layer. Note
similar air chord size distributions independently of the step that eq. [10] is homogeneous to a Darcy–Weisbach friction
height. Figure 7b presents the distributions of mean air factor. For the data shown in Figs. 5 and 6, the integration of
chord sizes at the same locations for the two step heights. eq. [9] along the step cavity yields an average friction factor

© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:09 PM
Color profile: Disabled
Composite Default screen

42 Can. J. Civ. Eng. Vol. 31, 2004

Fig. 7. Comparison of measured bubble chord sizes for two step heights h = 0.05 m and h = 0.10 m (dc/h = 1.7, (xl – xI)/dc = 7.5)
where I is inception f free-surface aeration: (a) bubble chord size probability distribution functions in the bubbly flow (C ~ 0.1), and
(b) distributions of mean bubble chord sizes for C < 0.5.

Table 4. Flow resistance in skimming flows. This result represents the average dimensionless shear
stress between the cavity flow and the main stream. In
Vertical step Dimensionless Darcy friction factor
monophase shear flow above a rectangular cavity,
height, h (m) flow rate, dc/h for air–water flow, fe
Wygnanski and Fiedler (1970) observed that the maximum
0.10 1.4 0.13 shear stress was almost independent from the distance from
1.5 0.10 the singularity, and their data yielded 8τmax/(ρVo2) = 0.18.
1.6 0.12 For cavity flows, Haugen and Dhanak (1966) and Kistler
1.7 0.15 and Tan (1967) observed similar results, which are of the
0.05 1.5 0.12 same order of magnitude as the present findings.
1.7 0.12 For the same flow rate as in Figs. 5 and 6, the flow resis-
2.0 0.11 tance estimate, derived from the friction slope, was fe = 0.15
2.2 0.10 (Table 4), where fe is the air–water flow friction factor. The
2.4 0.11 result is close to eq. [11]. Generally, present experimental
2.7 0.25 results demonstrated that the flow resistance was reasonably
3.2 0.16 well approximated by the integration of eq. [10] along step
Note: fe, air–water flow resistance deduced from the friction slope cavities. The finding is an improvement of the gross approx-
(Chanson et al. 2002). imation of Chanson et al. (2002), who assumed a constant
coefficient K.
For completeness, the “equivalent” friction factors derived
from the measured friction slope are presented in Table 4.
Lcav
1 8τmax The results are similar to the findings of Chanson and
[11]
Lcav ∫ ρ Vo2
dx s = 0. 34 Toombes (2002a) with a 22° slope and the analysis of Chan-
x =0
s son et al. (2002).
© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:10 PM
Color profile: Disabled
Composite Default screen

Gonzalez and Chanson 43

Summary and conclusion Chanson, H., and Toombes, L. 1997. Energy dissipation in stepped
waterway. Proceedings of the 27th IAHR Congress, San Fran-
An experimental investigation of skimming flow down a cisco, Calif., 10–15 August 1997. Edited by F.M. Holly, Jr., and
stepped chute was conducted in a large-sized facility (α = A. Alsaffar. Vol. D. pp. 595–600.
15.9°, W = 1 m) with two step heights (h = 0.1 and 0.05 m). Chanson, H., and Toombes, L. 2002a. Experimental investigations
The study focused on the air–water flow properties between of air entrainment in transition and skimming flows down a
step edges. Air–water velocity and turbulence measurements stepped chute. Canadian Journal of Civil Engineering, 29(1):
demonstrated a well-defined mixing layer developing down- 145–156.
stream of each step edge. In the developing shear layer, the Chanson, H., and Toombes, L. 2002b. Air–water flows down
velocity profiles had the same shape as classical monophase stepped chutes: turbulence and flow structure observations. In-
flow results (e.g., Tollmien and Goertler profiles), but the ternational Journal of Multiphase Flow, 27(11): 1737–1761.
rate of expansion of the mixing layer was greater, especially Chanson, H., and Toombes, L. 2002c. Energy dissipation and air
immediately downstream of the step edge. Maximum turbu- entrainment in a stepped storm waterway: an experimental
lent shear estimates in the shear layer yielded equivalent study. ASCE Journal of Irrigation and Drainage Engineering,
friction factors that were consistent with Darcy friction fac- 128(5): 305–315.
tors deduced from the measured friction slope. Overall, the Chanson, H., Yasuda, Y., and Ohtsu, I. 2000. Flow resistance in
skimming flow: a critical review. In Proceedings of the Interna-
findings confirmed the analogy between skimming flows and
tional Workshop on Hydraulics of Stepped Spillways, Zürich,
turbulent flows past cavities.
Switzerland, 22–24 March 2000. Edited by H.-E. Minor and
A comparative analysis of air–water flow properties for W.H. Hager. A.A. Balkema Publisher, Rotterdam, The Nether-
Froude similar flow conditions showed good agreement be- lands, pp. 95–102.
tween the two step heights in terms of dimensionless distri- Chanson, H., Yasuda, Y., and Ohtsu, I. 2002. Flow resistance in
butions of air content, velocity, and turbulence intensity, as skimming flows and its modelling. Canadian Journal of Civil
well as in terms of mean air content Cmean, dimensionless Engineering, 29(6): 809–819.
flow velocity Uw/Vc, and air–water flow velocity V90/Vc. Sig- Davies, P.O.A.L. 1966. Turbulence structure in free shear layers.
nificant differences were observed in terms of dimensionless AIAA Journal, 4(11): 1971–1978.
bubble count rates, turbulence levels, and bubble chord Djenidi, L, Anselmet, F., and Antonia, R.A. 1994. LDA measure-
sizes. The results highlighted some limitations of the Froude ments in a turbulent boundary layer over a D-type rough wall.
similitude for studies of skimming flows. Experiments in Fluids, 16: 323–329.
Elavarasan, R., Pearson, B.R., and Antonia, R.E. 1995. Visualiza-
tion of near wall region in a turbulent boundary layer over trans-
Acknowledgments verse square cavities with different spacing. Proceedings of the
12th Australasian Fluid Mechanics Conference, Sydney, Austra-
The writers acknowledge the assistance of Mr. Graham
lia, Vol. 1, pp. 485–488.
Illidge and Dr. L. Toombes. The first writer acknowledges
Haugen, H.L., and Dhanak, A.M. 1966. Momentum transfer in tur-
the financial support of the National Council for Science and bulent separated flow past a rectangular cavity. Journal of Ap-
Technology of Mexico (CONACYT). The second writer ac- plied Mechanics, Transactions American Society of Mechanical
knowledges the help of Mr. John LaBoon. Engineers, Sept.: 641–464.
Henderson, F.M. 1966. Open channel flow. MacMillan Company,
New York, N.Y.
References
Kistler, A.L., and Tan, F.C. 1967. Some properties of turbulent sep-
Boes, R.M. 2000. Zweiphasenstroömung und Energieumsetzung arated flows. Physics of Fluids, Part II, 10(9): S165-S173.
auf Grosskaskaden. Ph.D. thesis, VAW-ETH, Zürich, Switzer- Knight, D.W., and Macdonald, J.A. 1979. Hydraulic resistance of
land. artificial strip roughness. ASCE Journal of Hydraulic Division,
Chamani, M.R., and Rajaratnam, N. 1999. Characteristics of skim- 105(HY6): 675–690.
ming flow over stepped spillways. ASCE Journal of Hydraulic Liepmann, H.W., and Laufer, J. 1947. Investigation of free turbu-
Engineering, 125(4): 361–368. lent mixing. NACA Technical Note, No. 1257, August. National
Chanson, H. 1995. Hydraulic design of stepped cascades, channels, Advisory Committee for Aeronautics, Washington, D.C.
weirs and spillways. Pergamon, Oxford, U.K. Manso, P.A., and Schleiss, A.J. 2002. Stability of concrete macro-
Chanson, H. 1997. Air bubble entrainment in free-surface turbulent roughness linings for overflow protection of earth embankment
shear flows. Academic Press, London, U.K. dams. Canadian Journal of Civil Engineering, 29(5): 762–776.
Chanson, H. 1999. The hydraulics of open channel flows: an intro- Matos, J. 2000. Hydraulic design of stepped spillways over RCC
duction. Edward Arnold, London, U.K. dams. In Proceedings of the International Workshop on Hy-
Chanson, H. 2000. Forum article. Hydraulics of stepped spillways: draulics of Stepped Spillways, Zürich, Switzerland, 22–24
current status. ASCE Journal of Hydraulic Engineering, 126(9): March 2000. Edited by H.-E. Minor and W.H. Hager. A.A.
636–637. Balkema Publisher, Rotterdam, The Netherlands, pp. 187–
Chanson, H. 2001. The hydraulics of stepped chutes and spillways. 194.
A.A. Balkema, Lisse, The Netherlands. Matos, J., Yasuda, Y., and Chanson, H. 2001. Interaction between
Chanson, H., and Brattberg, T. 1998. Air entrainment by two- free-surface aeration and cavity recirculation in skimming flows
dimensional plunging jets: the impingement region and the very- down stepped chutes. Proceedings of the 29th IAHR Congress,
near flow field. Proceedings of the 1998 American Society of Me- Beijing, China, Theme D, Vol. 2. Edited by G. Li. Tsinghua
chanical Engineers Fluids Engineering Conference, FEDSM’98, University Press, Beijing, pp. 611–617.
Washington, D.C., 21–25 June 1998. American Society of Civil Minor, H.E., and Hager, W.H. (Editors). 2000. In Proceedings of
Engineers, New York, N.Y. Paper FEDSM98-4806. CD-ROM. the International Workshop on Hydraulics of Stepped Spillways,
© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:10 PM
Color profile: Disabled
Composite Default screen

44 Can. J. Civ. Eng. Vol. 31, 2004

Zürich, Switzerland, 22–24 March 2000. A.A. Balkema Pub- Cmean depth averaged air concentration defined as
lisher, Rotterdam, The Netherlands. (1 – Y90) Cmean = d
Ohtsu, I., and Yasuda, Y. (Editors). 1998. Hydraulic characteristics Do dimensionless diffusivity term
of stepped channel flows. Workshop on Flow Characteristics de equivalent depth of clear water (m) defined as
around Hydraulic Structures and River Environment, 13 Novem- Y90
ber 1998. University Research Center, Nihon University, Tokyo, de = ∫ (1 − C )dy
Japan. 0
Ohtsu, I., Yasuda, Y., and Takahashi, M. 2000. Characteristics of dj jet flow thickness (m)
skimming flow over stepped spillways. Discussion. ASCE Jour- dab air bubble size (m)
nal of Hydraulic Engineering, 126(11): 869–871. dc critical flow depth (m)
Rajaratnam, N. 1976. Turbulent jets. Development in water sci- F bubble count rate, i.e., the number of bubbles detected
ence, 5. Elsevier Scientific, New York, N.Y. by the probe sensor per second (Hz)
Rajaratnam, N. 1990. Skimming flow in stepped spillways. ASCE Fr Froude number defined as Fr = V / gde
Journal of Hydraulic Engineering, 116(4): 587–591. Discussion:
118(1): 111–114. f Darcy friction factor
Schlichting, H. 1979. Boundary layer theory. 7th ed. McGraw-Hill, fe Darcy friction factor for air–water flow
New York, N.Y. g gravity constant (m/s2) or acceleration of gravity
Sumer, B.M., Cokgor, S., and Fredsoe, J. 2001. Suction removal of H total head (m)
sediment from between armor blocks. ASCE Journal of Hydrau- h vertical height of steps (m)
lic Engineering, 127(4): 293–306. K inverse of the spreading rate of a turbulent shear layer
Sunyach, M., and Mathieu, J. 1969. Zone de mélange d’un jet plan. K′ integration constant
Fluctuations induites dans le cone à potentiel-intermittence. In- ks′ surface (skin) roughness height (m)
ternational Journal of Heat and Mass Transfer, 12: 1679–1697. L chute length (m)
Tantirige, S.C., Iribarne, A.P., Ojhas, M., and Trass, O. 1994. The Lcav step cavity length (m)
turbulent boundary layer over single V-shaped cavities. Interna- l horizontal length of steps (m) (measured perpendicular
tional Journal of Heat Mass Transfer, 37: 2261–2271. to the vertical direction)
Toombes, L. 2002. Experimental study of air–water flow properties lm Prandtl mixing length (m)
on low-gradient stepped cascades. Ph.D. thesis, Department of q discharge per unit width (m2/s)
Civil Engineering, The University of Queensland, Brisbane, Sf friction slope, Sf = –∂H/ ∂x
Australia. Tu turbulence intensity, Tu = u′/V
Toombes, L., and Chanson, H. 2000. Air–water flow and gas trans- Uw equivalent clear water flow velocity (m/s)
fer at aeration cascades: a comparative study of smooth and u′ root mean square of longitudinal component of turbulent
stepped chutes. In Proceedings of the International Workshop on velocity (m/s)
Hydraulics of Stepped Spillways, Zürich, Switzerland, 22–24 V velocity (m/s)
March 2000. Edited by H.-E. Minor and W.H. Hager. A.A. Vc critical flow velocity (m/s)
Balkema Publisher, Rotterdam, The Netherlands, pp. 77–84. Vo free-stream velocity (m/s)
Townes, H.W., and Sabersky, R.H. 1966. Experiments on the flow V90 characteristic velocity (m/s) where C = 0.90
over a rough surface. International Journal of Heat and Mass
v′ root mean square of lateral component of turbulent ve-
Transfer, 9: 729–738.
locity (m/s)
Tozzi, M., Taniguchi, E., and Ota, J. 1998. Air concentration in
W channel width (m)
flows over stepped spillways. Proceedings of the 1998 American
xl longitudinal distance measured in the flow direction (m)
Society of Mechanical Engineers Fluids Engineering Confer-
ence, FEDSM’98, Washington, D.C., 21–25 June 1998. Ameri- xs streamwise distance measured from the step edge (m)
can Society of Civil Engineers, New York, N.Y. Paper Y90 characteristic depth where the air concentration is
FEDSM98-5053, CD-ROM. 90% (m)
Wood, I.R. 1991. Air entrainment in free-surface flows. IAHR hy- y distance from the pseudo-bottom (formed by the step
draulic structures design manual No. 4. Hydraulic design con- edges) measured perpendicular to the flow direction (m)
siderations. A.A. Balkema Publisher, Rotterdam, The y50 distance normal to the invert where V = Vo /2 (m)
Netherlands. α channel slope
Wygnanski, I., and Fiedler, H.E. 1970. The two-dimensional mix- ∆x probe tip separation in the streamwise direction (m)
ing region. Journal of Fluid Mechanics, (Part 2), 41: 327–361. Ø dimensionless term, Ø = y/(ax)
µ dynamic viscosity (N·s/m2)
νt turbulent kinematic viscosity (m2/s)
List of symbols ρ density (kg/m3)
σ surface tension between air and water (N/m)
a empirical constant τ shear stress (Pa)
C air concentration defined as the volume of air per unit τmax maximum shear stress (Pa) in a shear layer
volume, also called void fraction τo average bottom shear stress (Pa)

© 2004 NRC Canada

I:\cjce\cjce3101\L03-066.vp
January 7, 2004 3:14:10 PM

Vous aimerez peut-être aussi