Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

ISH Journal of Hydraulic Engineering

ISSN: 0971-5010 (Print) 2164-3040 (Online) Journal homepage: https://www.tandfonline.com/loi/tish20

Aeration efficiency of bottom-inlet aerators for


spillways

M. Cihan Aydin

To cite this article: M. Cihan Aydin (2018) Aeration efficiency of bottom-inlet aerators for spillways,
ISH Journal of Hydraulic Engineering, 24:3, 330-336, DOI: 10.1080/09715010.2017.1381576

To link to this article: https://doi.org/10.1080/09715010.2017.1381576

Published online: 03 Oct 2017.

Submit your article to this journal

Article views: 122

View related articles

View Crossmark data

Citing articles: 1 View citing articles

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


https://www.tandfonline.com/action/journalInformation?journalCode=tish20
ISH Journal of HydraulIc EngInEErIng
2018, Vol. 24, No. 3, 330–336
https://doi.org/10.1080/09715010.2017.1381576

Aeration efficiency of bottom-inlet aerators for spillways


M. Cihan Aydin
Department of Civil Engineering, Bitlis Eren University, Bitlis, Turkey

ABSTRACT ARTICLE HISTORY


The bottom-inlet aerator can efficiently protect a spillway chute surface from cavitation damage due Received 6 February 2017
to high-flow velocities. This aerator, unlike conventional ones, is especially preferred at low Froude Accepted 15 September 2017
numbers and with wide spillway chutes to overcome submerged aerator conditions. In this study, the
KEYWORDS
aeration performance of bottom-inlet aerators was investigated for different Froude numbers and Air entrainment; bottom-
ramp heights. Some practical formulas based on experimental and numerical data were obtained to inlet aerator; cavitation;
estimate the air entrainment coefficient (and thus the average air concentration) of the bottom-inlet spillway
aerators. The findings were compared to some empirical results in published literature for conventional
aerators, and then discussed.

1. Introduction concentration rather than the bottom air concentration. Pfister


and Hager (2010a) also investigated the air transport devel-
Cavitation on the surface of the base of the spillway chute
opment along the downstream flow to optimize the hydraulic
exposed to high-velocity flow can produce serious damage to
design of the aerator. Kermani et al. (2015) used a method
the concrete surface, and thus the spillway chute may become
based on the nearest-neighbor model to forecast the damage
dysfunctional over time. For example, major cavitation damage
due to cavitation on spillway chutes, and tested their model
was observed on the spillway chutes of the Keban (Turkey) and
compared to the prototype data from the Shahid Abbaspour
Karun (Iran) dams between 1970 and 1980 (Pfister and Hager
dam spillway. They noted that their model gave a close estima-
2010a). Then, spillway aerators began to be used to prevent cav-
tion to observed damage in the prototype. Teng et al. (2016)
itation damage, especially for the concrete chutes subjected to
performed a 2D CFD (computational fluid dynamics) simu-
high-velocity flows. In a conventional aerator, the high-velocity
lation with a two-phase model to estimate the aerated flow
water flow over the spillway chute is deflected by a step and/
characteristics downstream of a spillway aerator, based on their
or ramp to create a cavity region under the jet that produces
experimental observations. They found that relatively good
a negative gauge pressure (sub-atmospheric pressure). The air
agreement was achieved with their experimental observation
produced by this sub-pressure region mixes into the chute
in the cavity, while numerical air concentrations were higher
flow along the stream direction. This self-aeration mecha-
than the experimental data in the far-field zone. Teng and Yang
nism is very effective in preventing cavitation in spillway flows.
(2016) carried out a 3D numerical simulation to determine
However, if the cavity sub-pressures under the jet drop down
and improve the aerating performance of the Bergeforsen dam
too low, the aerator may be submerged, which considerably
spillway, which is 35 m wide. They obtained reasonable agree-
reduces the aeration efficiency. Conventional aerators may also
ment between numerical and experimental data in terms of
become submerged in very wide chute conditions. In this case,
discharge capacity. They noted that the empirical relationships
the conventional aerator will be insufficient to ensure uniform
were insufficient to estimate the air entrainment for very wide
air concentration across the chute, and therefore the middle
spillway chutes, and the air distribution in the cavity must be
part of the chute is not protected from cavitation damage. These
guaranteed. Bhosekar et al. (2008) emphasized the need for an
submerged conditions appear, especially for low Froude num-
aerator, and its design and limitations for an orifice spillway.
bers. However, even at low Froude numbers, Chanson (1995)
Jothiprakash et al. (2015) also numerically simulated the flow
found that the submerged aerators might generate cavitation.
over an orifice spillway aerator. They used a 1:25 scale physical
Pinto et al. (1982) described a submerged aerator, and stated
model to calibrate their numerical model, described by differ-
that this aerator condition restricts air entrainment and should
ent characteristics of the air-water flow.
be prevented. Volkart and Ruschmann (1991) pointed out that
Ozturk et al. (2008) proposed a new aerator design called
a well-designed aerator device should ensure a uniform air dis-
the ‘bottom-inlet aerator’ to overcome submerged aerator con-
tribution across the chute width and should not be submerged.
ditions and to ensure better aeration along the chute width.
Some remarkable investigations have been presented in
They used air ducts arranged at a certain diameter and spacing
recent literature. Pfister and Hager (2010b) focused on the flow
along the chute width. In this way, the aerator can provide a
structure and the air transport phenomenon downstream from
sufficient and uniform air concentration along the chute width,
the chute aerator using a hydraulic model. The authors found
even for wide spillways and low Froude numbers. Laboratory
that the aerators primarily have an effect on the average air
tests were carried out to show the aeration performance of the

CONTACT  M. Cihan Aydin  mcaydin@gmail.com


© 2017 Indian Society for Hydraulics
ISH JOURNAL OF HYDRAULIC ENGINEERING 331

bottom-inlet aerators (Aydin and Aydin 2009). In this study, with dimensions is shown in Figure 1. Some displays from the
based on the experimental and numerical analyses performed experimental runs and the measuring devices used in the tests
by the author (Aydin 2005), the bottom inlet-aerators were are shown in Figures 2 and 3, respectively.
evaluated in respect to aeration performance, and the results The physical model tests were operated according to Froude
were compared to published literature. law similarity regardless of viscous and surface tension force.
Scale effects are expected for an air-water flow model in labora-
tory conditions, since Reynolds and Weber similarity cannot be
2.  Experimental tests
achieved together with Froude similarity. According to several
The laboratory tests were conducted at the Hydraulic Laboratory researchers, if the Reynolds number is greater than 105, air
of Firat University in Turkey. The experimental set-up was con- entrainment is not affected by the scale of the Froude model
structed using a plexiglas flume supported with a steel frame. In (Pinto et al. 1982). Reynolds numbers between 8.16  ×  104 and
order to provide a high-speed water jet – and therefore a high 1.73  ×  105 for the main flow in the chute were considered in the
Froude number – a jet box connected to a high-altitude reser- hydraulic model tests. Therefore, scale effects will be expected
voir with a constant level was used. The test flume was 4.00 m in the experimental observations. Because of the absence of
long and 0.50 m wide with a bottom angle of α = 15°. The flow prototype data for bottom inlet aerators, some relationships
discharge was controlled by a valve and measured by a flow were used in the literature to eliminate the scale effects. Ozturk
meter. An adjustable gate was installed at the interface between et al. (2008) used computational fluid dynamics (CFD) with
the jet box and the flume to allow for flow depth change. The the representative prototype dimensions of the bottom inlet
jet box and the adjustable gate assured the variable flow depths aerator for Reynolds numbers from 7.92  ×  106 to 7.43  ×  107. In
and Froude numbers upstream from the aerator. Five air ducts considering these Reynolds numbers, the Froude similarity is
of 12 mm in diameter were arranged across the bottom of the required at a maximum scale of 1/57. Therefore, it is noted that
channel at 10  cm intervals. The air velocities in the supply the scales used in laboratory cause considerable scale effects.
ducts were measured with a digital velocity meter. Three sets In order to eliminate the scale effects on air entrainment rate
of experiments were conducted in which the ramp heights, tr, in the model values, Equation (1) –first presented by Kokpinar
were fixed at 4 , 6  and 8 mm with a ramp angle of θ = 7.59°. and Gogus (2002) with different coefficients – was used in the
The range of Froude numbers tested was from 4.27 to 9.75, over following form based on the numerical data with the prototype
which significant aeration occurred. The experimental set-up size in this study.

Figure 1. Demonstration of experimental setup (dimensions are in cm).


332 M. CIHAN AYDIN

Figure 2. Views of experimental runs.

Figure 3. Digital measuring devices using in the hydraulic model tests.

where β and βm are the prototype and model values of the


𝛽 = 4.407(𝛽m )1.264 (1) air entrainment rates, respectively. Based on prototype data,
a similar equation with coefficients of 5.194 and 1.150 for
ISH JOURNAL OF HYDRAULIC ENGINEERING 333

symmetrical aeration was also used by Kokpinar and Gogus relationship between the air entrainment rate and the Froude
(2002). number for different ramp heights. The Froude number and the
ramp height have important effects on the air entrainment. The
variations in the air entrainment rate with the Froude number
3.  Aeration efficiency
and the relative ramp height are illustrated in Figure 4(a) and
It is known from published literature that the air entrainment (b), respectively, for different ramp heights. As seen in Figure
rate for conventional aerators is mainly a function of the 4(a), while the effect of the ramp height on the air entrainment
upstream Froude number. The air entrainment rate related to is considerably small compared to the Froude number effects,
the following non-dimensional parameters as a result of the its increasing effect on the air entrainment can be clearly seen
dimensional analysis for a bottom-inlet aerator (Ozturk et al. in Figure 4(b).
2008) is: The relationships between the values for β and Fo numbers
are given in Equation (3) with the scale effects and in Equation
(4) without the scale effects. Both equations were obtained by
( )
t D2
𝛽 = f Fo , r , , tan 𝛼 (2)
h Lh regression analysis with a correlation of 93%, and limited by
ranges of 4.27 ≤ Fo ≤ 9.75 and 0.07 ≤ tr/h ≤ 0.41.
where β is the air entrainment rate – defined as a ratio of the
air discharge supplied by the aerator to the water discharge
𝛽m = 0.017(Fo − 2) (3)
in the flume – i.e. Qa/Qw. Fo is the upstream Froude number
defined as the ratio of the inertia forces to the gravity forces:
Fo = Uw/(gh)1/2, Uw is the approaching flow velocity (in m/s),
𝛽 = 0.047(Fo − 2.80) (4)
g is gravitational acceleration (in m/s2), h is the approaching Variations of the air entrainment rate for the model and pro-
flow depth, D is the diameter of the aeration holes and L is the totype, and fitted linear lines with Equations (3) and (4) are
distance between consecutive air holes. The non-dimensional shown in Figure 5.
parameter, D2/Lh, represents the ratio of the air hole area with Ozturk et al. (2008) suggested an equation to estimate the
a diameter of D to the flow cross-sectional area supplied by the air entrainment rate of the bottom-inlet aerators with prototype
hole. The experimental setup was only operated to observe the dimensions as a function of Fo, D2/Lh, and tr/h (R2  =  98.8%),
neglecting the effects of the chute slope on the air entrain-
ment, which is quite low. Based on numerical data, a revised
equation (Equation 5) was determined for even lower Froude
numbers with R2  =  98.6% (Fo  <  2.80 approximately). Equation
(5) assumes that β is linearly related to the Froude number.

) D2 0.20 tr 0.35
( ) ( )
(5)
(
𝛽c = 0.14 Fo − 2.8
Lh h

where subscript c indicates computational results. Equation


(5) may be given with the following limitations on the rel-
evant parameters: 2.80  ≤  Fo  ≤  10.00, 0.17  ≤  tanα  ≤  0.57,
0.04 ≤ tr/h ≤ 0.50, 0.032 ≤ D2/Lh ≤ 0.539 and θ = 7.59°.
In Figure 6, the air entrainment rates calculated from
Equation (5) were compared to the model and the represent-
ative prototype data by Equation (1). This figure shows that
while the representative prototype data converted by Equation
(1) are consistent with the results of Equation (5), the physi-
cal model data including scale effects are not. Thus, it is sug-
gested that Equation (5) can be used to estimate the required
air entrainment rate ensured by the bottom-inlet aerators for
preventing cavitation damage, especially in concept project
stages – perhaps before model studies. Note that, as mentioned
by Pfister and Hager (2010a, 2010b), aerators mainly increase
the average air concentration, while their effect on the bottom
is low. Therefore, the air distributions, the flow depth and the
detrainment rate downstream of the aerator should be also
considered for effective protection of the chute bottom.

4.  Jet length


It is known that the jet length over a ramp is a basic parameter
of the air entrainment of the aerators, and specifies the general
air transport zone. The jet length, Lj, is the length between the
ramp end and the impact point of the flow jet where the pres-
Figure 4.  Variations of the air entrainment rate with (a) Froude number, (b) sure is a maximum. Several empirical relations for conventional
Relative ramp height. aerators have been suggested to estimate relative jet length for
334 M. CIHAN AYDIN

where Aa and Aw are the cross-sectional areas of the aerator


duct and flow respectively.
[√ ]
Lj 1.5 tr
= 0.77Fo (1 + sin 𝛼) + Fo tan 𝜃 for 0 < Lj ∕h < 50
h h
(7)

In this study, to determine relative jet length for the bot-


tom-inlet aerators, a new relationship was obtained based on
the numerical data with prototype dimensions (R2 =  0.86) as
follows:

Lj ( )0.21 ( )0.16
D2 tr
= 2.29Fo (1 + tan 𝛼)0.22 (8)
h Lh h

In Figure 7, the relative jet length for the bottom-inlet aerator


Figure 5. Distribution of air entrainment rate with Froude number. was compared to that of the conventional aerators proposed
by Pfister (2008), and Kokpinar and Gogus (2002). As seen in
this figure, while Equations (6) and (7) gave consistent results,
Equation (8) for bottom-inlet aerators gave slightly lower and
linear results – unlike conventional ones. The reason for this is
probably the sub-atmospheric pressure in the cavity under the
jet; in contrast, the other equations were derived in jet cavities
at atmospheric pressures.
Kokpinar and Gogus (2002) presented the following rela-
tionship to estimate the air entrainment rate through the lower
jet as a function of the relative jet length, the relative aerator
area (Aa/Aw) and the bottom slope. To eliminate scale effects
in Equation (9), they used Equation (10) for a symmetrical
aerator (Kokpinar and Gogus, 2002).
( L )0.83 [( ) ]0.24
j Aa
𝛽m = 0.0189 (1 + tan 𝛼) (9)
h Aw

Figure 6. Comparison of representative prototype and model data.


𝛽 = 5.194𝛽m1.150 (10)
Pfister and Hager (2010b), based on Froude number and ramp
no offset in published literature, e.g. by Kokpinar and Gogus
angle, also proposed an empirical relationship to estimate β
(2002) with Equation (6), and Pfister and Hager (2010a) with
with R2  =  0.93 as follows:
Equation (7), as follows:
𝛽 = 0.0028F2o 1 + Fo × tan 𝜃 − 0.1 (11)
[ ]
( )0.44 [
Aa −0.087
(L ) ]
j 0.22 tr
( )1.75
= 0.28 Fo (1 + 𝜃) (1 + tan 𝛼) (6)
h m h Aw For the bottom-inlet aerators, as shown in Figure 8, a signifi-
cant correlation was determined between β and the relative jet
length Lj/h. When the effects of the Froude number are con-
sidered to be linear, the following relationship can be defined
to estimate β based on Lj/h for the bottom-inlet aerators with
a high correlation of R2  =  0.98.
( )0.92
) Lj
(12)
(
𝛽 = 0.006 Fo − 2.8
h
5.  Air concentration
According to Peterka (1953), the average air concentration
should be 0.01 <  Ca  <  0.06 to avoid cavitation damage. The
average air concentration can be easily calculated using Ca = β/
(1  +  β). However, Pfister and Hager (2010b) outlined that
the definition of β was not enough, and that the streamwise
development of air transport along the downstream flow of
the aerator should also be known to predict cavitation dam-
age. The following equations, to determine the average and
Figure 7. Variations of relative jet length with Froude number. bottom air concentration along the stream, based on extensive
ISH JOURNAL OF HYDRAULIC ENGINEERING 335

where Cai is the incipient air concentration (Cai  ≈  0.10). Figure


9(a) shows the distributions of average air concentration vs. Fo
for different approaches. Note that in Figure 9(a), the data for
the experiments, and the representative prototype (Equation
1) and Equation (5) are obtained for the bottom-inlet aerator.
On the other hand, the results of Kokpinar and Goguş (2002),
and Pfister and Hager (2010b), are calculated from Equations
(9–10) and Equation (11), respectively, for conventional aer-
ators. While the results of Equation (5) are quite compatible
with that of Kokpinar and Gogus (2002), the Pfister and Hager
(2010b) equation gives slightly higher values. The experimen-
tal values are considerably lower than the others due to scale
effects. Figure 9(b) illustrates the streamwise air concentration
downstream from the aerator at x/Lj = 3 and x/Lj = 9. It is seen
that the air concentration from x = 3Lj to 9Lj is sufficient to
overcome cavitation damage when considering Peterka’s (1953)
limitations.
Figure 8.  Correlation between the air entrainment rate and the relative jet
length.
6. Conclusions
experimental investigations, were suggested by Pfister and The following conclusions are outlined based on this study:
Hager (2010b):
(1) The air entrainment is mainly related to the upstream
Lj Froude number Fo. However, the other dimension-
Ca(3Lj ) = 0.008 + Cai , for 5 ≤ Lj ∕h ≤ 40, (13)
h less parameters influencing the air entrainment of
bottom-inlet aerators are the relative ramp height
x
( )
Ca = Ca(3Lj) + 0.02 − 3 sin (𝛼 − 30◦ ), for 3 ≤ x∕Lj ≤ 9 tr/h, the ratio of D2/Lh and also the channel bed slope
L (tan α), but its effect is relatively low. The air entrain-
(14)
ment is also non-linearly related to the relative jet
length Lj/h.
(2) Although Pfister and Hager (2010b) concluded that
the minimum Fo for efficient air entrainment in con-
ventional aerators is approximately Fo  =  5 for θ  =  6°,
the bottom-inlet aerators may be more effective even
for lower Froude numbers, e.g. Fo   <   5. This aera-
tor also provides uniform air concentration along
the chute width, especially for wide spillway chutes
because each partial width L of the chute is supplied
by an air duct with diameter D arranged along the
chute width.
(3) Equations (5) or (12) can be used to estimate the air
entrainment rate of bottom-inlet aerators, especially
at the initial stage of a project.
(4) The findings of this study are compatible with those
in published literature for conventional aerators.
(5) The bottom-inlet aerator ensures enough air concen-
tration (Ca   >   0.08) to overcome cavitation damage
even at x   =    3Lj and x   =    9Lj downstream from the
aerator – in terms of average air concentration.

Acknowledgments
This work was supported by Firat University Scientific Research Projects
Unit (Project No: FUBAP-1034).

Disclosure statement
No potential conflict of interest was reported by the author.

Funding
Figure 9. Average air concentration vs. Fo: (a) Comparison of different approach, This work was supported by the Firat Üniversitesi (FUBAP-1034).
(b) Streamwise air concentration at x/Lj = 3 and 9.
336 M. CIHAN AYDIN

References Peterka, A.J. (1953). “The effect of entrained air on cavitation pitting.”
Proc., IAHR-ASCE Joint Conf. Int. Hydraul. Convent., 1953, 507–518.
Aydin, M.C. (2005). “CFD Analysis of Bottom-Inlet Spillway Aerators.” Pfister, M. (2008). “Schussrinnenbelüfter: Lufttransport ausgelöst durch
PhD Thesis (in Turkish), Firat University Graduate School of Natural interne Abflussstruktur [Chute aerators: Air transport due to internal
and Applied Sciences, Department of Civil Engineering, Elazığ. flow structure].” Mitteilung, 203, H.-E. Minor, ed., Laboratory of
Aydin, M.C., and Aydin, S. (2009). “An experimental study on bottom- Hydraulics, Hydrology and Glaciology (VAW), ETH, Zurich, (in
inlet aerators (in Turkish).” IVth National Water Engineering Symposium, German).
6–10 July 2009, 37–42, Istanbul. Pfister, M., and Hager, W.H. (2010a). “Chute aerators II: hydraulic design.”
Bhosekar, V.V., Jothiprakash, V., and Deolalikar, P.B. (2008). “Hydraulics J. Hydraul. Eng., 136, 360–367.
of aerator for office Spillway.” ISH J. Hydraul. Eng., 14(1), 28–40. Pfister, M., and Hager, W.H. (2010b). “Chute aerators I: transport
Chanson, H. (1995). “Predicting the filling of ventilated cavities behind characteristics.” J. Hydraul. Eng., 136, 360–367.
aerators.” J. Hydraul. Res. Int. Assoc. Hydraul. Res., 33(1995), 361–372. Pinto, N.L.S., Neidert, S.H., and Ota, J.J.. (1982). “Aeration at high velocity
Fadaei Kermani, E., Barani, G.A., and Ghaeini-Hessaroeyeh, M. (2015). flows.” Water Power Dam Constr., 34(2), 34–38.
“Prediction of cavitation damage on spillway using K-nearest neighbor Teng, P., and Yang, J. (2016). “CFD modeling of two-phase flow of a
modeling.” Water Sci. Technol., 71(3), 347–352. spillway chute aerator of large width.” J. Appl. Water Eng. Res., 4(2),
Jothiprakash, V., Bhosekar, V.V., and Deolalikar, P.B. (2015). “Flow 163–177.
characteristics of orifice spillway aerator: numerical model studies.” Teng, P., Yang, J., and Pfister, M. (2016). “Studies of Two-phase flow at a
ISH J. Hydraul. Eng., 21(2), 216–230. chute aerator with experiments and CFD modelling.” Model. Simul.
Kökpınar, M.A., and Göğüş, M. (2002). “High-speed jet flows over Eng., Article ID 4729128, 1–11.
spillway aerators.” Can. J. Civ. Eng., 29(6), 885–898. Volkart, P., and Rutschmann, P. (1991). “Aerators on spillways.” Air
Ozturk, M., Aydin, M.C., and Aydin, S. (2008). “Damage limitation- a new entrainment in free surface flows, I.R. Wood, ed., A.A. Balkema,
spillway aerator.” Int Water Power Dam Constr., 60(5), 36–40. Rotterdam, 85–114.

You might also like