Vous êtes sur la page 1sur 20

Journal of Hydraulic Research

ISSN: 0022-1686 (Print) 1814-2079 (Online) Journal homepage: https://www.tandfonline.com/loi/tjhr20

Study of air entrainment and aeration devices

H. Chanson

To cite this article: H. Chanson (1989) Study of air entrainment and aeration devices, Journal of
Hydraulic Research, 27:3, 301-319, DOI: 10.1080/00221688909499166

To link to this article: https://doi.org/10.1080/00221688909499166

Published online: 19 Jan 2010.

Submit your article to this journal

Article views: 180

View related articles

Citing articles: 51 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tjhr20
Study of air entrainment and aeration devices
Etude de l'entraïnement d'air et des aérateurs
H. CHANSON
Engineering Consultant,
143, rue de la Pompe, 75116 Paris, France

SUMMARY
Cavitation erosion damage to spillway surfaces may be prevented with the use of aeration devices (aerators)
introducing air in the layers close to the channel bottom in order to reduce cavitation damage.
The performance of a spillway aerator are studied on steep spillway model with high velocities (from 4 m/s up
to 15 m/s) and new air concentration and velocity measurements were performed. The data are presented
and analysed developping new equations. The behaviour of the air demand relationship between the air
discharge and the subpressure in the cavity beneath the nappe is analyzed. All these results provide a better
understanding of the air entrainment processes above a spillway aerator and the quantity of air entrained can
be obtained.
Experimental data at the impact point of the jet are shown. The results indicate a strong deaeration process
occuring in the impact region and reducing the effect of the aeration occuring above the aerator.

RESUME
Les dommages causes par l'érosion par cavitation sur les coursiers des évacuateurs de crues peuvent être
empêchés a Paide d'aérateurs, introduisant artificiellement de Pair dans Pécoulement proche de la surface
des coursiers.
Les performances des aérateurs ont été étudiées sur un modèle a forte pente avec grandes vitesses (de 4 m/s a
15 m/s), et de nouveaux mesurements de concentration en air et de vitesse ont été obtenus. Les résultats sont
présentés et analyses avec de nouvelles equations. Le comportement de la demande en air, entre Ie débit en
air et la pression dans la cavilé située sous le jet, est analyse a partir de nouvelles mesures. L'ensemble de ces
résultats permet une meilleure comprehension des phénomènes d'entraïnement d'air, et la quantité d'air
entramée est calculée.
Les mesures obtenues au point d'impact du jet sont présentées. Les résultats montrent la presence d'un
processus de déaération dans cette region, réduisant les effets de faération de Pécoulement au-dessus de
Paérateur.

1 Introduction

1.1 Presentation

The irregularities on the spillway surfaces will in a high speed flow cause small areas of flow
separation and in these regions the pressure will be lowered. If the velocities are high enough the
pressure may fall to below the local vapor pressure of the water and vapor bubbles will form.
When these are carried downstream into high pressure region the bubble collapses giving rise to
high pressures and possible cavitation damage. Experimental investigations show that the
damage can start at clear water velocities of between 12 to 15 m/s and up to velocities of 20 m/s
it may be possible to protect the surfaces by streamlining the boundaries, improving the surface
finishes or using resistant materials (Volkart and Rutschmann 1984) [1].

Revision received February 6, 1989. Open for discussion till December 31, 1989.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 27. 1989, NO. 3 301


C(%)

X 13.0 MPa
6- O 18.8 MPa

_<D

X D
2- Volume
at loss (cm3)

0 10 20 30 40

Fig. 1. Relation between the volume loss and the air concentration V = 46 m/s.Russell and Sheenan (1974).
Relation entre l'érosion (en volume) et la concentration en air I7 = 46 m/s Russell et Sheenan
(1974).

Peterka (1953) [2] and Russell and Sheehan (1974) [3] performed experiments on concrete
specimens and showed that air concentrations of 1-2% reduce substantially the cavitation erosion
and above 5-7% no erosion was observed. The results are plotted in Fig. 1 where C is the average
air concentration of the flow at atmospheric pressure and ambient temperature. The tests were
performed over two hours in each case with concretes of different compressive strengths and the
flow velocity at the points where cavitation occurred was 46 m/s.
The entrained air through the free surface of the flow may protect the spillway floor from cavita­
tion damage if the free-surface aeration process provides a sufficient air concentration near the
bottom (i.e. C > 7%). If there is not enough surface aeration (i.e. downstream of a gate) or if the
tolerances of surface finish required to avoid cavitation are too severe (i.e. V > 30 m/s), air can be
artificially introduced by devices called aerators and located on the spillway floor and sometimes
on the side walls.

1.2 Aerator Devices

A small deflection in a spillway structure (i.e. ramp, offset) tends to deflect the high velocity flow
away from the spillway surface (Fig. 2). In the cavity formed below the nappe, a local subpressure

Patm AP

Fig. 2. Aerator device.


Aérateur de fond.

302 JOURNAL DE RECHERCHES HYDRAUL1QUES, VOL. 27, 1989, NO. 3


beneath the nappe (A/5) is produced by which air is sucked into the flow (Qa"rlct).
Vischer et al. (1982) [4] detailed the properties of the different types of aeration devices. Usually a
combination of a ramp (slope 0, length Z-r:,mp). an offset (height fs) and a groove provides the best
design: the ramp dominates operation at small discharges while the groove provides space for the
air supply and the offset enlarges the trajectory of the jet at higher discharges.
Volkart and Chervet (1983) [5] have studied on model the behaviour of a large range of aerator
and Volkart and Rutschmann (1984) present several examples of air supply system.

1.3 Definitions

The local air concentration C is defined as the volume of air per unit volume and this will normal­
ly be taken as a time averaged value. We define the characteristic depth d as:

d= \ (1 -C) *dv (1)


o
wherey is measured perpendicular to the spillway surface. It is also necessary to define a charac­
teristic depth for self aerated flow. For both model and prototype measurements a depth that is
well defined and convenient is that where the average air concentration is 90% (Y%).
A depth averaged mean air concentration for the flowing fluid can be defined from:
(]-CnKJ*Yw = <i (2)

An average water velocity is then:

t/w = — (3)
d
The water discharge per unit width may be transformed:
<7w = ( l - C r o e a n ) * £ / w * K90 (4)
A characteristic water velocity (V90) is defined as that at Y90.

1.4 Experiments

The author Chanson (1988) [6] performed experiments on a 1 : 15 scale model of the Clyde dam
spillway with a slope a = 52.33°. The model provided Froude numbers in the range 3 to 25 with
initial average flow velocities from 3 m/s to 14 m/s. By adjusting the gate at the entry of the flume
the initial flow depth was from 20 mm to 120 mm. The first aerator configuration included a ramp
of 5.7° (30 mm height, 300 mm length) and an offset of 30 mm height. This geometry was the same
as that used by Low (1986) [7]. The second configuration had no ramp and an offset of 30 mm
height as that used by Tan (1984) [8].
New conductivity probes were developed to record air concentration measurements with a single
tip probe and velocity measurements of air-water mixture using a two-tips velocity probe and a
cross-correlation method.
In the first part the mechanisms of air entrainment above an aerator are described. Then the
experimental results obtained on spillway model are presented and discussed.

JOURNAL OF HYDRAULIC' RESEARCH, VOL. 27, 1989. NO. 3 303


2 The mechanisms of air entrainment

2.1 Presentation

The regions of flow above a bottom aerator are on a long spillway (Fig. 3): 1. the approach flow
region, 2. the transition zone, 3. the aeration zone, 4. the impact point region, 5. the downstream
flow region and 6. the equilibrium region.
The approach zone may be in a region where some of the surface is aerated. The transition zone
coincides with the length of the ramp. The deflector changes the mean perpendicular pressure
field and increases the shear stress on the spillway floor. This change alters the turbulent field and
these have a strong influence on the lower air-water interface in the aeration zone.
Without a ramp there is still a pressure change at the lip of the aerator from a hydrostatic pressure
distribution to a negative pressure gradient. Indeed both with and without the ramp the pressure
on the upper and lower free surfaces at the lip are respectively atmospheric and the cavity
pressure. The rise velocity of a spherical air bubble (diameter db) subject to such a negative
pressure gradient becomes a fall velocity Chanson (1988):

4 * g * db
ul 3*C,
PN * cos a + (5)

where z is the vertical axis and PN the pressure gradient number defined as

AP
PN = -■
Qw * g * a
In the aeration region air is entrained by high intensity turbulent eddies close to the air-water
interfaces and this type of aeration called nappe entrainment occurs on both the lower and upper
free surfaces of the jet Ervine and Falvey (1987) [9].
The bottom pressure attains its maximum at the impact point of the jet (Fig. 4). The rollers at
the rear of the cavity entrain an additional quantity of air by plunging jet entrainment. The flow

Approach flow region


/ Transition zone Patm Impact
point

Aeration zone Deaeration

Cavity
Impact point atm - P
region
Rollers'

Downstream flow kp
k / 4 ^
region
X

Equilibrium - P
flow region
Pressure at the spillway bottom

Fig. 3. Flow regions above an aerator. Fig. 4. Impact point region.


Ecoulement au-dessus d'un aérateur de fond. Point d'impact du jet.

304 JOURNAL DE RECHERCHES 11YDRAUL1QUES. VOL. 27, 1989, NO. 3


is highly turbulent and a high energy loss occurs in the impact region. Variations of the position
of the impact point occur and it is suggested that these are caused by subpressure oscillations
which occur in the cavity below the jet.
The impact region is characterized by a rapid air concentration redistribution with a de-aeration
process occurring immediately downstream of the point of impact (Fig. 4). At the end of the jet
the flow is subject to a rapid change of pressure distribution from a negative pressure gradient to a
maximum pressure gradient at the impact point (higher than the hydrostatic pressure gradient)
and finally to the hydrostatic pressure distribution far downstream. The high pressure gradient in
the impact region may cause the strong de-aeration observed at the free surface.
The downstream flow region includes a rapidly varied flow region and a gradually varied flow
region. This ends when the air concentration and velocity distributions reach the equilibrium
flow region far downstream of the aerator.

2.2 Discussion
The study of air entrainment on spillway aerator is complex because of the interaction between
the different air entrainment processes and the author showed that the air entrainment (Fig. 5)
above an aerator is characterized by: 1. air entrainment through the upper and lower free surfaces
of the water jet called nappe entrainment, 2. plunging jet entrainment at the intersection of the
water jet and the rollers, and 3. air recirculation in the cavity below the jet.
Let consider the control volume ABCDEF in the flow region above the aerator (Fig. 6). We
define: q"^' the net air entrainment through the upper free surface BC, q{°™" the net air entrain­
ment through the lower air-water interface AF, gj'"" 8 the plunging jet entrainment, #a!rclrc the air
recirculation and Qa"rle' the air discharge supplied by the air inlets.
The continuity equation for the air phase in the control volume ABCDEF of the flow is:

where q$ is the initial quantity of air entrained within the flow through AB and qf® is the
quantity of air entrained within the flow through CD.
The continuity equation applied to the cavity below the jet is:

^alr i „recirc _ lower , plung /T-,


Tjr ~ Hair Hair ' Hair \'t
W
The combination of the equations [6] and [7] yields:

q^-q™ = qTe' + ~ (8)


w
This equation suggests that the air entrainment above an aerator is only function of the air
discharge provided by the air supply system and the net air entrainment through the upper air-
water interface of the jet. These parameters are function of the fluid properties, the spillway and
aerator geometry, the geometry of the air inlets, the flow properties, and the undernappe cavity
properties.
The air discharge supplied by the air inlets is usually studied as a function of the subpressure
in the cavity and the flow conditions for a given aerator geometry. This relationship is called the
air demand of the aerator and will be detailed later. The subpressure in the cavity is obtained by
combining the air demand and the pressure loss characteristic of the air supply system. Hence the

JOURNAL OF HYDRAULIC RESEARCH, VOL. 27, 1989, NO. 3 305


Air entrainment
through the
upper interface

Air entrainment
through the
lower interface

Air
recirculation
Plunging
jet '
entrainment
Impact
point

Fig. 5. Air entrainment above an aerator.


Entraïnement d'air au-dessus d'un aérateur.

upper
lair

lower
qair

recrrc
lair
plung -' D\T
lair '' *N
Impact
point

Fig. 6. Control volume ABCDEF.


Volume de controle ABCDEF.

306 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 27, 1989, NO. 3


jet trajectory calculations will provide: 1. the position of the impact point, 2. the geometry of the
cavity beneath the nappe and 3. the angle of the jet with the spillway surface.
We are not able to measure the net quantity of air entrained through the upper free-surface of the
jet. However in the aeration region a method to estimate the quantity of air entrained within the
flow from air concentration and velocity measurements will be presented later and experimental
results obtained on spillway model will be shown. Hence the equations [6] and [7] should be used
rather than the equation [8]. The interactions between the air discharge Q'a"'c\ the plunging jet
entrainment <7a!runE, the air recirculation g™"rc, the lower nappe entrainment q]™e' and the upper
nappe entrainment ql?/" are unknown.
The processes occurring in the impact region remain unclear. It must be emphasized that the air
concentration and velocity measurements are not accurate because of the high turbulence and
rapid redistributions of air concentration and velocity occurring in this region. Hence it is not
possible to estimate the air exchange through the free-surface of the jet in the impact region.

3 The air demand study


3.1 Introduction
The air demand is defined as the relationship between the air discharge provided by the air supply
system, the subpressure in the cavity and its distribution along the nappe, and the flow charac­
teristics.
According to Kobus (1984) [10], Pinto (1984) [11], Laali and Michel (1984) [12], the Reynolds and
Weber numbers do not influence the air demand in any significant way if: 1. the Reynolds
number is greater than 10' and
V
2. the Weber number Wx = is greater than 400.

The ratio W7, is defined by Pinto (1984) using the jet length as characteristic length. Pinto at al.
(1982) [13] performed experiments on a series of hydraulic model whose scale varied from 1 :8
through to 1 : 50 and were able to show that the model reproduced the prototype air demand for
all water discharges for scales larger than 1:15. For scales 1 :30 and 1 : 50 the correct air demand
was only reproduced for the larger discharges. For a large scale model and for a given aerator
configuration the dimensionless air demand becomes:
Qirilet
rhl = ~-=f(Fr,Tu,PN)

where Fr is the Froude number, Tu the turbulence intensity and FN the pressure gradient number.
There is little information available on the effects of turbulence and often the studies neglect the
influence of the turbulence intensity. On the spillway model, the turbulence intensity distribu­
tions are almost constant for a given aerator geometry and a given flow depth. Thus we may
consider Tu as a parameter of the spillway which depends on the geometry and we then get:
/? inlet =/(Fr,/» N ) (9)

JOURNAL OF HYDRAULIC RESEARCH. VOL. 27. 1989. NO. 3 307


3.5-1 3.5-
D Fr = 2190 P
• Fr = 2Z30 3.0-
+ Fr =21.67
O Fr = 21.21
Fr = 20.28 2.5-
A +++
A Fr =17.32
M Fr = 14.93 2.0- • d/ts = 0.77
+ dAs = 1.15
*t

■ Fr =13.33
d/ts = 1.65
1.5- f A
A d/ts =2.17
X d/ts = 2.70
1.0-
■ d/ts = 3.67

0.5-
x"
X, Fr
0.0 -r~ — -
r -1
10 15 20 25 30

Fig. 7. Air demand for <///^ = 0.95 aerator with Fig. 8. Air demand for P\=0 aerator with
ramp (0 = 5.7°). ramp (0 = 5.7°).
Demande en air pour <7//s = 0.95 Demande en air pour PN = 0.
aérateur equip d'une rampe (0 = 5.7°). Aérateur équip d'une rampe (0 = 5.7°).

3.2 Results

For different flow depths the experimental results may presented as: 1. /? mlct = ƒ (PN) for different
Froude numbers (Fig. 7) and 2. /? m l e t =ƒ(ƒ>) for different pressure gradient numbers (Fig. 8).
The relationship between the three dimensionless numbers /?' nlet , / r a n d /°N shows that a decrease
of air discharge brings an increase of the nappe subpressure and for the same nappe subpressure
the air discharge increases with the Froude number. Fig. 8 indicates a change of mechanism of
air entrainment for Fr~ 5. At low Froude numbers (Fr< 5) the air demand is less affected by the
cavity subpressure and most of the air is entrained by plungingjet entrainment at the rear of the
cavity below the jet.

4 The Aeration Region

4.1 Air Concentration

4.1.1 P r e s e n t a t i o n
The aeration region characterizes the flow region from the lip of the aerator to the vicinity of the
rollers where the flow is a two-dimensional jet subject to a negative pressure gradient between the
upper and lower nappes. For high Froude numbers a large quantity of air is entrained through
both the upper and lower free surfaces. This does not occur at low Froude number (i.e. Fr< 5) if
the conditions of nappe entrainment are not reached. For axi-symmetric water jets discharging
horizontally, Ervine and Falvey (1987) suggest that an estimate of the condition for the onset
of free surface aeration is:

0.275
(10)
Tu

308 JOURNAL OF RECHFRCUHS HYDRAULIQUES. VOL. 27, 1989, NO. 3


where Tu is the turbulence intensity. If this condition is reached, aeration occurs and the air is
entrained by high intensity turbulent eddies close to the air-water interfaces.
At the end of the deflector a solid inner jet core of clear water (C = 0 %) is observed and it is
reduced along the channel while the flow is aerated through the free surfaces. When the aeration
reaches the central part of the jet (Fig. 9), that becomes aerated. If the waterjet is long enough, a
fully-aerated jet region starts developing downstream of the point where the central part of the jet
becomes aerated and each section is then characterized by its minimum air concentration C rain
(Fig. 10) which increases along the jet as the quantity of air entrained augments.

4.1.2 D i s c u s s i o n
In the free surface aeration the existence of a non-aerated water core prevents any interaction
between the air entrainment processes at the lower and upper interfaces. Consider a small control
volume the continuity equation for the air is:

I)
C = — div </~*r (ID
Dt

w h e r e q.Mr is t h e air flux:

<7air = — D * V C + C * IIT (12)

C is the air concentration, D the diffusivity, ur the rise velocity of bubble subject to a pressure
gradient (equation (5)) and we assume that the density of air is constant. The continuity equation
may be developed at each air-water interface assuming 1. a steady flow, and

dC dC
dx dy

Consider the upper free-surface. Assuming 1. a homogeneous turbulence and 2. an uniform flow
velocity t/w, the continuity equation becomes:

dC dC ., d2C
Uy, * — + u, * cos 6 * — = D * — Y (13)
dx dy dy

0.9 1

Fig. 9. Flow above a bottom aerator aeration Fig. 10. Air concentration distribution in the
region. aeration region Fr=\9.2, L/d,t= 10.1
Eeoulement au-dessus d'un aérateur de Profil de concentration en air dans la
fond zone d'aération. zone d'aération Fr— 19.2, L\dn= 10.1

JOURNAL Ot- HYDRAULIC RESEARCH. VOL. 27. 1989. NO. 3 309


where Dl° is the turbulent diffusivity at the upper interface, ur is the rise velocity, x is the axis
parallel to the streamline,y is the axis perpendicular to the streamline and 0 is the angle between
the streamline x and the horizontal. For

— * cos 8 <s: 1,
£/»

the solution is a Gaussian distribution:

y
C = erf (14)
* X * 1+ * COS 0 * —
t/w \ Um x
where the function erf is defined as Spiegel (1974) [14]:

-ti/2
>rf{u) ] d/
l/2
At the lower air-water interface the rapid change of shear stress is dominant and if 1. the rise
velocity term is small and 2. the turbulence is homogeneous the continuity equation becomes:

dC dU d / dC
u * —+c — = —[D * — (15)
dx dx dy \ dy

where U is the velocity, D is the turbulent diffusivity at the lower interface, x is the axis along the
streamline and ƒ the axis perpendicular to the streamline. The effect of the removal of the shear
stress at the lower interface is to allow the fluid to accelerate and at the end of the deflector the
term C * dUjdx is large.

C=l

x =x+ -JY^*COS6 * y
c

C = 0.5
uu
C=0
Water jet
streamline

C=l
Fig. 11. Air concentration distribution in the free-surface aeration region w r <0.
Profil de concentration en air dans la zone d'aération w r <0.

310 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 27, 1989, NO. 3


The equation (15) may be rewritten:

dC d I. n, dC
U * — = — \(D + D(>) * —
dx dy \ dy

where D is a term related to the longitudinal velocity gradient and defined as:
.+dv
n dC dU
D°* — = I C * — * dy
dv y OX

For the boundary conditions shown on Fig. 11 and for a gradually varied parameter D° the
solution of the continuity equation is a Gaussian distribution:

C = 1 - erf (16)
2 * *x
U

At the lip of the aerator the rapid reduction ofshear stress at the lower interface allow the fluid to
accelerate and the effect of the acceleration term C * dUjdx is important. As we move down­
stream this term decreases to zero and when

dU d2C
C « ; /)
~d7
the continuity equation (15) becomes a classical diffusion equation:

dC d2C
i/w* D
dx dy1

and for the above boundary conditions (Fig. 11) the solution becomes:

v
C=\- erf -
i *— *x

0.5

V -~£
0.0 -i—i—i—i—i—i—i—i—■—i
b. 0.0 0.2 0.4 0.6 0.8 1.0
Fig. 12. Comparison of air concentration distributions with Gaussian profiles.
a. fr=14.9, rf0 = 34.8 mm
b. f>=5.98, f/„ = 80.7 mm
Comparaison des profils de concentration en air avec une distribution Gaussienne.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 27, 1989, NO. 3 311


In the free-surface aeration region the solutions of the continuity equation for both the upper and
lower interfaces (equations (14) and (16)) indicate that the air concentration distribution follows
a Gaussian distribution.
For different flow conditions the data were compared with Gaussian distributions computed
from K90 and YH) (Chanson 1988) and typical results are presented in Figs. 12.
From the air concentration measurements obtained in the free-surface aeration it is possible to
estimate of the turbulent diffusivity at the upper free surface (Z)u = D {equation (14)}) and the
diffusivity at the lower air-water interface (DL = D 4- D° {equation (16)}). For low subpressures
(i.e. PK < 0.01) the results are presented in Table 1.
The diffusivity at the lower free surface averages the effect of the turbulence and the velocity
gradient. The results show that the diffusivity at the lower interface is almost 10 times larger than
the turbulent diffusivity at the upper free-surface.
In the fully aerated jet region the inner core of the jet becomes aerated and the air concentration
distributions may no longer be fitted by a Gaussian curve.

Table 1. Diffusivity at the lower and upper interfaces for / J N <0.01


Difïusivité aux interfaces supérieure et inférieure pour P N <0.01
do/t, Z)u DL
(mm) (m2/s) (m2/s)
0.77 0.08- KT4 15.4-10 4
1.15 1.0510 4 11.710 4
2.70 0.63-10 4 3.6410 4

4.2 The quantity of air entrained

4.2.1 Definition
The quantity of air entrained within the flow is defined over a cross-section as:

1 Gair(y) * KM * «ty
0
<?air=
fair

For the flow on spillway several authors Cain (1978) [15], Wood (1985) [16] indicate that the slip
ratio K = V.^\V^ is almost equal to 1 between 0 and 90% of air concentration. Assuming that the
variations of the air density across the section are small, the quantity of air entrained may be then
rewritten in term of air concentration as:
Y9O

qm= j C * V *dy (17)


0

If the water discharge qv can be estimated as:


Y90
</ w = i ( 1 - C ) * V *d.v (18)
0

and assuming:
Y.)u

£/ w * Yw= j |/*d.v (19)


0

312 JOURNAL DE R E C H E R C H E S 1IYDRALILIQUES, VOL. 27, 1989, NO. 3


the equation (17) may be rewritten:

<7air = £ 4 * I ' w - ' / w

Therefore the dimensionless quantity of air entrained /? = </air/c/w becomes a function of the
characteristic depths d and K,0 of the considered section:

(20)

For the author's experiments where the air concentration and the velocity measurements were
performed, it was possible to calculate the experimental value of <?ail. (equation (17)) and

I V * d.v.
0

Figs. 13 and 14 present the comparison between the equations (19) and (20) with the experimental
results. The graph 14 shows that the equation (20) provides a good estimate of the quantity of air
entrained and the Fig. 13 shows that the assumptions (18) and (19) give reasonable results.
It is worth noting that the exact definition of the water discharge qw is:

0w * <?w = i ö w M * Ki.v) * d>'

Between y = 0 and y = Y90 the slip ratio is equal to 1 and hence the above definition becomes:
00

y, 0 J 8wG>) * Vv(y) * d.v


<7w = i (1 - C) * V * Ay + —
0 Qw
It must be remarked that the second term cannot be measured and hence the best estimation of
the water discharge <?w becomes:
'90

<7w= i ( 1 - C ) * V*dy

U.8- Uw*Y90
m3/s/m S^ +
0.6- y / \

0.4- 4*r

0.2-
}V*dy (data)
m3/s/m
0.0- 1 1
1 1 1 ' 1
0.0 0.2 0.4 0.6 0.4

Fig. 13. Comparison between the equation (19) Fig. 14. Comparison between the equation (20)
and ihe experimental results. and the experimental results.
Comparaison entre l'équation (19) et les Comparaison entre l'équation (20) et les
mesures expérimentales. mesures expérimentales.

JOURNAL OF H Y D R A U L I C RESEARCH, VOL. 27. 1989, NO. 3 313


Fig. 15. Comparison between the equation (22) and the experimental data.
Comparaison entre Péquation (22) et les mesures experimental.

The equation (20) is an important result because it shows that the quantity of air entrained may be
obtained from the air concentration measurements only.

Application to a free jet


For the flow above the cavity (Fig. 9) the total quantity of air entrained within the flow is defined
between the lower and the upper free-surfaces of the jet as:
vnipper
1
90
Total
Vair j C * V *dy (21)
1
On

where y9u0pper and Y^,w" are the characteristic depths defined for C = 90% at the upper and lower
air-water interfaces (Fig. 10).
With assumptions similar to the equations (18) and (19) the equation (21) then becomes:
Total yupper ylower
o Total _ 4;.ir _ ^90 ~ ^90 _ ] .^

From the air concentration and velocity measurements performed above the aerator, the calcula­
tions from the experimental data of the quantity of air entrained within the flow computed from
the equation (21) is compared with the equation (22).
The results (Fig. 15) are within the accuracy of the data and this suggests that the assumptions are
reasonable and that the equation (22) is a reasonable estimate of the quantity of air entrained
within the flow.
The air concentration profile (Fig. 10) gives a minimum in the inner core of the jet and at each
position in the aeration region we can define a minimum air concentration Cmm and a charac­
teristic depth Yc where C = Cmin. In the aeration region we define the quantity of air entrained
within the upper region of the flow (YCv < v < KU )as:
yupper
1
un

,u C * V * dv
Y-

314 JOURNAL DE RECHERCHES 11YDRAULIQUES, VOL. 27. 1989, NO. 3


The dimensionless quantity of air entrained within the upper region /? l becomes:
yuppcr
1
90
i (1 - C) * di-
ii y/uppcr y v
CLl ÏW rC
?U _ " _ ~ min 'C,„
F (23)
<7w c/ d

Identically we can define the quantity of air entrained within the lower region of the flow
(K,!;rr<>'<nmJas:
Yc
*- mm
<7air — 1 C * V * d\'
Y j.„,

A dimensionless quantity of air entrained within the lower region of the flow jiL is defined and
turns into:
Y
cmir,
I (1 - C) * Ay
,,L Y ylower wlower

4.2.2 R e s u l t s
From the author's experiments the quantity of air entrained within the flow may be deduced.
Typical curves are plotted in Figs. 16 where L is the distance form the end of the deflector and d0
the initial flow depth. The computed position of the impact point is shown.
The graph 16 shows that the quantity of air entrained within the upper region of the jet
(Yc . < y< Ygop") increases substantially when the pressure gradient increases and this may be
caused by the effects of the pressure gradient on the rise bubble velocity equation (5).
The experimental results on spillway model Chanson (1988) indicate that the increase of the
quantity of air entrained within the lower region of the jet is almost independent of the Froude

i.u - Impac t point


+ B Total
■ BU
0.8-
• C
0.6- l> t . ^ ^ D

0.4-
* B
0.2-
A .
Ud
o.o- ' i ' ' ' ' i i—i—i—i—i—i

50 100 150 b. o
Fig. 16. Quantity of air entrained for rf0=22.9 mm and /r=19.5.
a. / T l c ' = 0.679, />N = 0.011
b. /T" c l = 0.0, />N = 0.742
Quantité d'air entraïne pour rf0=22.9 mm el Fr= 19.5.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 27, 1989, NO. 3 315


number and the pressure gradient. These suggest that the rapid change of shear stress at the lower
interface is dominant. The interaction ofthe air recirculation in the cavity (of finite volume) and
the effect o f t h e pressure gradient on the bubble velocity is a complex phenomenon.
Globally the maximum quantity of air entrained within the flow (point C) increases when the
pressure in the cavity becomes increasingly below atmospheric and this may caused by a greater
nappe entrainment through the upper air-water interface. The curvature of the jet is more
pronounced for high pressure gradient and the higher pressures at the impact point due to a
greater angle between the jet and the spillway surface cause a stronger local de-aeration in the
impact region.

5 The impact region

In the impact region the flow is subject to a rapid change of pressure distribution from a negative
pressure gradient to a maximum pressure gradient at the impact point (higher than the hydro­
static pressure gradient) and a strong de-aeration process occurs (Fig. 16).
The results obtained on the Clyde dam spillway model indicate that the flow conditions
(reference depth d„ mean air concentration CJ at the end of the impact region are almost
independent ofthe flow discharge, the subpressure AP in the cavity and the air flow provided by
the air supply system Qjfrlet. These conditions are only function ofthe depth of water c/u in the
approach flow region ofthe aerator (Table 2) and the position ofthe start ofthe downstream flow
region is function ofthe position ofthe impact point of the jet. The latter is calculated from the jet
trajectory equations, where the subpressure AP in the cavity is obtained from the air demand
through the air inlets.
It must emphasized that the air concentration and velocity measurements in this region are not
accurate because o f t h e high turbulence and the rapid redistributions of air concentration and
velocity.

Table 2. Flow conditions at the end ofthe impact region on the Clyde dam spillway model, offset height:
f,=30 mm
Conditions d'écoulement la fin de la zone d'impact pour Ie modèle reduit de l'évacuateur de crues
du barrage de Clyde
initial mean air reference
depth concentration depth slope
dalts C, a conditions reference
0.77 0.32 0.87 52.33° no ramp Chanson (1988)
1.15 0.26 1.10 52.33° no ramp Chanson (1988)
2.70 0.12 2.10 52.33° no ramp Chanson (1988)
1.67 0.29 1.15 51.30° ramp ^ = 5.7° Low (1986)

6 Conclusion

The air entrainment above an aerator is characterized by a nappe entrainment through the upper
and lower free surface of the jet, an additional quantity of air entrained by plunging jet entrain­
ment and a recirculation process within the cavity.
A method has been developed to compute the quantity of air entrained within the How from the
air concentration distributions. This analysis provides simple equations which have been verified
from air concentration and velocity distribution.

316 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 27, 1989, NO. 3


The study of air entrainment is strongly related to the jet trajectory calculations and the air
demand study. For given flow conditions and aerator geometry the air demand study provides:
1. the relationship Qi" =f(AP) between the subpressure beneath the nappe and the air dis­
charge supplied by the air inlets;
2. for a given air duct configuration the operating point characterized by AP and Qa"rlet. From the
determination of the subpressure in the cavity the jet trajectory calculations provide;
3. the volume of the cavity beneath the nappe;
4. the position of the impact point;
5. the angle of the jet with the spillway floor at the impact point.
The experimental results suggest a strong de-aeration process occurring in the impact region but
new instrumentation must be developed to obtain more accurate measurements in this region.

7 Acknowledgements

The author wishes to thank the Civil Engineering Department, University of Canterbury
(New Zealand), the University Grant Committee (New Zealand) and the Ministry of Works and
Development (New Zealand) for their financial support, and Professor I. R. Wood who super­
vised this project.

Notation

AA duct area below the surface spillway (m 2 )


C air concentration defined as the volume of air per unit volume
Cd drag coefficient
C mm minimum air concentration in a cross section above the aerator

d characteristic depth (m) defined as: d = j (1 — C) * Ay where y is measured


perpendicular to the spillway surface
db air bubble diameter (m)
d0 characteristic depth in the approach flow region
V
Fr Froude number defined as: Fr = =
Vg * d
g gravity constant (m/s 2 )
local value: g = 9.8050 m/s 2 (Christchurch, New Zealand)
Kair
K slip ratio defined as the ratio of the air velocity over the water velocity K =

L distance along the spillway from the end of the deflector (m)
Ljet distance of the impact point of the jet from the end of the deflector (m)
AP
PN pressure gradient number defined as: PN =
Ow * g * d
Qair air discharge (nrVs)
Qa"rlcl air discharge provided by the air supply system (mVs/m)
Qw water discharge (nrVs)
0W * V * d
Re Reynolds number defined as: Re =
/"
is offset height (m)

JOURNAL OF HYDRAULIC RESEARCH, VOL. 27, 1989, NO. 3 317


w
Tu turbulence intensity defined as: Tu = —

Uv average water velocity (m/s) defined as: C/w = —


d
ur rise bubble velocity (m/s)
u root mean square of axial component of turbulent velocity (m/s)
V velocity (m/s)
K9n characteristic velocity at Y90 (m/s)
W channel width (m)
x distance from the end of the deflector (m)
y depth measured perpendicular to the spillway surface (m)
Yc . characteristic depth (m) where C = C min
Kio characteristic depth (m) where the air concentration is 10%
y,'ower characteristic depth (m) where the air concentration is 10% at the upper free surface
of the jet above the aerator
y1u0pf":r characteristic depth (m) where the air concentration is 10% at the lower free surface
of the jet above the aerator
Y9a characteristic depth (m) where the air concentration is 90%
J^9ower characteristic depth (m) where the air concentration is 90% at the upper free surface
of the jet above the aerator
u p
Y9 tf " characteristic depth (m) where the air concentration is 90% at the lower free surface
of the jet above the aerator
a spillway slope
P dimensionless quantity of air entrained within the flow
/? inlet dimensionless air discharge provided by the air supply system
L
/? dimensionless quantity of air entrained within the lower region of the jet in the aeration
region
/jTotai dimensionless total quantity of air entrained within the flow in the aeration region
/JL' dimensionless quantity of air entrained within the upper region of the jet in the aeration
region
AP difference between the pressure above the flow and the air pressure beneath the nappe
(Pa): AP = P0- />cavity
<p angle between the ramp and the spillway
H dynamic viscosity of water (N • s/m 2 )
0 angle between the water jet and the horizontal
g air density of air (kg/m 3 )
QW density of water (kg/m 3 )

References / Bibliographie

1. VOLKART, P. and RUTSCHMANN, P., Air Entrainment Devices, Mitteilungen der Versuchsanstait fur
Wasserbau, Hydrologie und Glaziologie, No. 72, Zurich, Switzerland, 1984.
2. PETERKA, A. J., The Effect of Entrained Air on Cavitation Pitting, Joint meeting paper, IAHR/ASCE,
Minneapolis, USA, Aug. 1953.
3. RUSSELL, S. O. and SHEENAN, G. J., Effect of Entrained Air on Cavitation Damage, Canadian Journal or
Civil Engineering, Vol. 1, 1974.

318 JOURNAL DE RECHERCHES HYDRAUL1QUES, VOL. 27. 1989, NO. 3


4. VISCHER, D., VOLKART, P. and SIGENTHALER, A., Hydraulic Modelling of Air Slots on Open Chute Spill­
ways, Int. Conf. on Hydraulic Modelling, BHRA Fluid Engineering, Coventry, England, September
1982.
5. VOLKART, P. and CHERVET, A., Air Slots for Flow Aeration. Mitteilungen der Versuchsanstalt fur
Wasserbau, Hydrologie und Glaziologie, No. 66, Zurich, Switzerland, 1983.
6. CHANSON, H., Study of Air Entrainment and Aeration Devices on Spillway Model, Research Report 8-88,
University of Canterbury, New Zealand, October 1988.
7. Low, H. S., Model Studies of Clyde Dam Spillway Aerators, Master Report, Ref. 86-6. University of
Canterbury, New Zealand 1986.
8. TAN, T. P., Model Studies of Aerators on Spillway, Master Report, Ref. 84-6, University of Canterbury,
New Zealand 1984.
9. ERVINE, D. A. and FALVEY, H. T., Behaviour of Turbulent Water Jets in the Atmosphere and in Plunge
Pools, Proc. Instn Civ. Engrs, Part 2, 1987, 83, Mar., pp. 295-314.
10. KOBUS, H., Local Air Entrainment and Detrainment. Symposium on Scale Effects in Modelling Hy­
draulic Structures, IARH, 1984.
11. PINTO, N. L. DE S., Model Evaluation of Aerators in Shooting Flow, Symposium on Scale Effects in
Modelling Hydraulic Structures, IARH, 1984.
12. LAALI, A. R. and MICHEL, J. M., Air Entrainment in Ventilated Cavities: Case of the Fully Developed
"Half-Cavity", Trans, of ASME, J. of Fluids Engineering, September 1984, Vol. 106, pp. 327-335.
13. PINTO, N. L. DE S., NEIDERT, S. H. and OTA, J. J., Aeration at High Velocity Flows, Water Power& Dam
Construction, Feb.-Mar. 1982.
14. SPIEGEL, M. R., Formules et Tables de Mathematiques, Mathematical Handbook of Formulas and
Tables, McGraw-Hill Inc., New York, USA, 1974.
15. CAIN, P., Measurements within Self-Aerated Flow on a Large Spillway, Ph.D. Thesis, Ref. 78-18, Univer­
sity of Canterbury, Christchurch, New-Zealand, 1978.
16. WOOD, I. R., Air Water Flows, 21st Congress IAHR, Aug. 1985, Melbourne, Australia.

JOURNAL OF HYDRAULIC RESEARCH. VOL. 27. 1989. NO. 3 319

Vous aimerez peut-être aussi