You are on page 1of 84

Air entrainment with plunging jets

Experimental study about air entrainment with free


overfall jets from circular channels and air bubble
intake with submersible pumps in sewer sumps

M.Sc. thesis of
Arnout Smit

WL | Delft Hydraulics

Delft University of Technology


Faculty of Civil Engineering and Geosciences
Section Environmental Fluid Mechanics
2
Preface
This master thesis is an experimental study to air entrainment in sewer sumps. The
work has been carried out within the framework of the faculty Civil Engineering and
Geosciences at Delft University of Technology.

I would like to thank my graduation committee for their assistance and comments;
WL|Delft Hydraulics for the opportunity to work on this subject and building the
physical model set-up; ITT Flygt B.V. and A.B Dietzel for lending the submersible
pump; at last I want to thank Nander van der Plicht and Rogier Smit for their help
with the report.

A. Smit

Graduation committee:
Prof.dr.ir. G.S. Stelling
Prof.dr.ir. F.H.L.R. Clemens
Dr.ir. W.S.J. Uijttewaal
Ir. C.L. Lubbers

March 2007

WL | Delft Hydraulics

Delft University of Technology


Faculty Civil Engineering and Geosciences
Section of Environmental Fluid Mechanics

3
Table of contents

List of symbols ............................................................................................................... 6


Summary........................................................................................................................ 8
Chapter 1: Introduction .............................................................................................. 10
Chapter 2: Literature study......................................................................................... 11
2.1 Introduction ................................................................................................................. 11
2.2 Dimensional analysis ................................................................................................... 11
2.3 Air entrainment mechanisms ..................................................................................... 13
2.3.1 Viscous and low velocity jets............................................................................................... 13
2.3.2 Jet surface instabilities ......................................................................................................... 14
2.3.3 Other entrainment mechanisms ............................................................................................ 15
2.3.4 New findings on entrainment mechanisms .......................................................................... 15
2.4 Basic Hydrodynamic features of liquid jets and bubbles ........................................ 15
2.4.1 Impact velocity of the jet ..................................................................................................... 15
2.4.2 Water jet diameter ................................................................................................................ 16
2.4.3 Jet surface roughness ........................................................................................................... 16
2.4.4 Break-up length of liquid jets............................................................................................... 17
2.4.5 Boundary layer of a surrounding gas ................................................................................... 18
2.4.6 Bubble rise velocity ............................................................................................................. 18
2.5 Minimum entrainment velocity ................................................................................. 18
2.6 Volumetric flow rate of entrained air........................................................................ 19
2.6.1 Methods of entrained gas measurements ............................................................................. 19
2.6.2 Air entrainment regions ....................................................................................................... 20
2.6.3 Effect angle of impact on air entrainment ............................................................................ 23
2.7 Entrainment ratio Qa/Qw ............................................................................................ 23
2.7.1 Entrainment ratio Qa/Qw for different jet regions ................................................................. 23
2.7.2 Effect of surface tension and liquid viscosity on entrainment ratio Qg/Ql ........................... 25
2.8 Characteristics of bubble dispersion ......................................................................... 26
2.8.1 Bubble size distribution ....................................................................................................... 26
2.8.2 Penetration depth of entrained bubbles ................................................................................ 27
2.8.3 Aeration and bubble zone .................................................................................................... 28
2.9 Free overfall with circular open channels ................................................................. 29
2.9.1 Introduction .......................................................................................................................... 29
2.9.2 End Depth Ratio he/hc .......................................................................................................... 30
2.9.3 End Depth Ratio with sloping channels ............................................................................... 30
2.9.4 Dimensionless discharge ...................................................................................................... 30
2.9.5 Properties of circular open-channels .................................................................................... 31
2.10 Conclusions literature study..................................................................................... 32
Chapter 3: Description of experimental model set-up and measurement program . 33
3.1 Model reservoir ........................................................................................................... 33
3.2 Horizontal circular open tube as approach flow channel ........................................ 35
3.3 Submersible sewer pump............................................................................................ 35
3.4 Types of flow regulation ............................................................................................. 36
3.5 Specific measurement attributes ................................................................................ 36

4
3.6 Mesurement program and procedures...................................................................... 39
3.6.1 Measurement program ......................................................................................................... 39
3.6.2 Measurement procedures ..................................................................................................... 40
3.6.3 Imposed parameters ............................................................................................................. 40
3.6.4 Measured parameters ........................................................................................................... 42
Chapter 4: Qualitative observations and discussion.................................................. 44
4.1 Observations on plunging jets and air entrainment ................................................ 44
4.1.1 Jet surface roughness ........................................................................................................... 44
4.1.2 Shape of the plunge point..................................................................................................... 47
4.1.3 Features which influence the jet shape during the fall ......................................................... 47
4.1.4 Plunge point enclosing funnels ............................................................................................ 49
4.1.5 Jet plunge angle.................................................................................................................... 50
4.1.6 Approach flow channel inclination ...................................................................................... 51
4.2 Observations on air bubbles and air bubble plumes ............................................... 51
4.2.1 Transport of air into the reservoir ........................................................................................ 51
4.2.2 Air bubble properties ........................................................................................................... 52
4.2.3 Submerged jet obstruction ................................................................................................... 53
4.2.4 Reservoir circulation ............................................................................................................ 54
4.3 Observations on pump air intake .............................................................................. 55
Chapter 5: Results and analysis ................................................................................. 57
5.1 End depth measurements ........................................................................................... 57
5.2 Horizontal fall distance ............................................................................................... 58
5.3 Circular jet cross section fall height .......................................................................... 60
5.4 Aeration length ............................................................................................................ 61
5.5 Penetration depth ........................................................................................................ 64
5.6 Air entrainment discharge ......................................................................................... 69
5.7 Pump impeller revolution rate ................................................................................... 72
5.8 Pump air discharge ..................................................................................................... 73
Chapter 6: Conclusions .............................................................................................. 78
Chapter 7: Recommendations .................................................................................... 80
References ................................................................................................................... 81
Appendix A: Schematic drawing of model set-up ...................................................... 83

5
List of symbols
Ae cross sectional area of the water stream at tube end (m)
Aj cross sectional area of plunge point (m2)
C constant
da air bubble diameter (m)
dN diameter of the nozzle (m)
dj diameter of the jet in plunge point used in literature (m)
dr reference diameter of the jet
dv volume-equivalent bubble diameter (m)
D diameter of the approach flow tube (m)
DH hydraulic diameter (m)
Ej kinetic energy per second of the jet (Nm/s)
 
Fr Froude number  = VN  (-)
 gd N 

hc critical depth (water depth at critical flow speed) (m)
ĥc dimensionless critical depth (= hc / D) (-)
he end depth (tube end water depth) (m)
ĥe dimensionless end depth (= he / D) (-)
hf funnel depth around the plunge point (m)
hm capillary head of model water (mm)
ht capillary head of tap water (mm)
H fall height is the vertical distance between the lower tube end and reservoir water level (m)
Hcir fall height when the jet cross section is circular shaped at a given discharge (m)
ˆ
H dimensionless circular fall height (= Hcir / D) (-)
cir
He fall height of mass centre of the tube stream (= H + 2/3 * he) (m)
Hp penetration depth (m)
Hu fall height when the jet cross section is U-shaped (m)
Hw water depth in reservoir (m)
lN length of cylindrical section of a nozzle (m)
La aeration zone (= Lb – Ljx) (m)
Lb bubble zone (m)
LB breakup length of the jet (m)
Lj length of the jet (m)
Ljx horizontal fall distance (m)
LR ratio between model and prototype length (-)
LT approach flow tube length (m)
Ltp horizontal distance between tube edge and vertical pump intake axis (m)
qw discharge per meter width of weir (m3/h/m)
Qa air entrainment discharge (l/min)
Qap air intake at the pump (l/min)
Qcir circular discharge is the water discharge when the plunge point is circular shaped at a given
jet fall height (l/s)
Qg gas discharge (m3/s)
Ql liquid discharge (m3/s)
Qw water discharge (= jet discharge = pump discharge) (l/s)
Qˆ w dimensionless discharge  = Qw  (-)
 g D 2.5 

Relength Reynolds number of the jet accompanying air boundary layer  = Ve L j  (-)
 
 νw 
S approach channel slope (-)
Sc critical slope (-)
t time (s)
Tj fall time (s)
turbulence intensity  = V j  (-)
'
Tu
 V 
 j 

6
ur bubble rise velocity (m/s)
ve horizontal flow speed at tube end (m/s)
vj horizontal flow speed of the jet at plunge point (m/s)
Va local velocity of boundary layer air of the jet (m/s)
Ve total flow speed at tube end (m/s)
Vj total flow speed of the jet at plunge point (m/s)
V j′ mean squared jet velocity fluctuation (m/s)
Vmin minimum entrainment velocity (m/s)
VR ratio between the model and prototype velocity (-)
we vertical flow speed at tube end (m/s)
we average vertical flow speed at tube end (m/s)
wj vertical flow speed of the jet at plunge point (m/s)
Weber number  = VN d j ρ a  (-)
2
Wea
 σ 

Weber number at the nozzle  = VN d N ρ w  (-)
2
WeN
 σ 
 µw  (-)
Z Ohnesorge number  = 
 d N ρ wσ 

αc critical jet plunge angle (degrees)
αe angle of the stream axis at the tube end (degrees)
αes angle of the stream surface at the tube end (degrees)
αj jet plunge angle (degrees)
δe angle end (see figure 2.16) (degrees)
θN nozzle contraction angle (degrees)
µw dynamic viscosity of water (Pa*s)
νw kinematic viscosity of water (= µ/ρ) (m2/s)
ρ density (kg/m3)
σ surface tension (N/m)
φ contact angle between jet surface and reservoir water surface

Subscripts
a air
ap air at pump
b bubble
B breakup
c critical flow
cir circular shaped jet cross section
e tube end
F nozzle feed
g gas
j jet
l liquid
M model
min minimum
N nozzle
p penetration
P prototype
r reference or rise
R ratio
s surface
t tap water
T tube
u U-shaped jet cross section
v volume
w water
x horizontal

7
Summary
Many pressurized sewer systems do not reach their design capacity discharge due to
enlarged resistance in the pipe system. An extensive investigation has brought to light
that air pockets in pipe systems are an important cause of high resistance in sewer
system pipe lines. Air intake by pumps in sumps is one of the main reasons of air
pocket formation in sewer systems. But there is still little known about air entrainment
due to plunging jets from circular open channels, and air intake by submersible pumps
near plunging jets, especially at the scale of real sewer sumps.
The research objective is to achieve knowledge and insight into air entrainment
from free overfall water jets from horizontal open channels and air bubble intake with
submersible sewer pumps.

Figure I: Simple schematic overview of research model

The main questions are:


• How does the air entrain into the reservoir?
• Where and how many air bubbles can be expected in a certain configuration
without any specific measures?
• What is the sphere of influence of the submersible sewer pump?
• What is a robust measure to prevent air coming into the sewer pump?

It is found in literature that jet surface disturbances are the main cause of the air
entrainment with plunging jets. These disturbances entrap air at the location between
the jet surface and the reservoir water surface when the surfaces do not fit exactly.
In this experimental research it was found out that the perimeter shape of the
intersection between the falling jet and the reservoir water surface (plunge point) is
also an important factor for the amount of air entrainment. For free overfall jets from
circular channels the horizontal jet cross section shape is at first U-shaped and
becomes along the fall gradually circular. The found dimensionless empirical formula
H Qw
for the circular fall height is: cir = 102.0
D g D 2.5

At first instance, a constant or increasing penetration depth was expected with


increasing jet fall heights, but it turned out that the opposite effect was more often the
case. The air/water entrainment ratio determines mostly the penetration depth and the
higher the entrainment ratio, the smaller the penetration depth. A reason for that is,
the entrained air bubbles break up the submerged jet and so the jet loses its
momentum.

8
The aeration length appeared to be independent of the plunging jet fall height. The
aeration length is mainly affected by the jet water discharge. The empirical maximum
aeration length with an exceedance frequency of once every eight to ten seconds is:
s
La ,max = 0.5Qw0.5 , with the dimension of constant 0.5 equal to
m

From experiments with pump air intake experiments the following observations
are made. Air bubbles are only taken in by the pump when they are in the direct
neighborhood of the pump inlet. The influence sphere of the pump is for discharges
Qw ≤ 15.0 not more than a radius of 0.10 m around the pump inlet centre.

An unexpected phenomenon is the effect of a pre-rotation at the pump inlet on the


pump air intake discharge. A pre-rotation is a circulation in front of the pump inlet as
a result of the pump impeller revolution rate that is not working in its Best Efficiency
Point (B.E.P.). The pre-rotation pushes away the air bubbles near the pump inlet and
complicate air from taking in by the pump. But this effect is only marginal with
impeller revolution rates twice the B.E.P. and is therefore not an effective and
economical way to prevent air bubbles to be taken in by the pump.

In this thesis a couple of mostly empirical formulas taken from literature are
compared with the results of the experiments. Almost all formulas proposed in
literature are only representative for small jets produced with nozzles. Only Dey
(1998, 2001) and Aigner (1999) have described some recent findings about free
overfall flows in circular channels. But they give no information about the falling jets
produced with the overfall channel flows, nor the air entrainment with the plunging
jets. The only formula from literature that turned out to be valid for the experiments
done is the dimensionless discharge formula of Rajaratnam & Muralidhar (1964):
11
Qw  h 6
= 1.54  e  .
g D 2.5 D

The measured results are obtained with a specific model set-up. Other similar
models predict different results, especially with the pump air intake experiments. The
measurements are very sensitive to all kinds of variations. Therefore, the results are
only useful for giving an indication of the in reality occurring orders of magnitudes.

The most robust and effective way to prevent air bubbles coming deep into the
sewer sump near the submersible pump inlet is to break up and bend the vertical
submerged jet in the reservoir. When the submerged jet hits a horizontal object below
the water surface, the high velocity flows in the reservoir are bended horizontal. Air
bubbles cannot be brought deeper into the reservoir and rise up unhindered in the
horizontal flows to the water surface. Because real sewer sumps have varying water
levels, the horizontal object in the reservoir should be below the lowest water level.

This master thesis is an experimental research about a specific subject which is


hardly addressed in literature. Because of that, this thesis is mainly a basic description
of all occurring phenomena with the used model set-up. Other research programs can
expand the results with more theory and better predictions. The experimental program
is easily expandable as well.

9
Chapter 1: Introduction
Many pressurized sewer systems do not reach their design capacity discharge due to
enlarged resistance in the pipe system. Because the design capacity discharge is not
reached, sewer overflows will work more often at heavy rainfall, which is not good
for the environment. As a result, more powerful pumps have to be installed and sewer
system pipes have to be enlarged. Older systems are not always able to withstand the
bigger pressures from these more powerful pumps which results in a higher safety
risk.
An extensive investigation has brought to light that air pockets in pipe systems are
an important cause of high resistance in sewer system pipe lines. How the air gets into
the sewer systems is unclear. Possible sources are rotting gasses, degassing of sewage
as a result of the lower pressures downstream, or air intake at the pump inlet by air
entrainment in sewer sumps.

The research program CAPWAT by WL|Delft Hydraulics and TUDelft, which


started in 2003 (Lubbers & Clemens, 2005, 2006), aims at the behavior and influence
of air bubbles on the capacity and the energy loss in pressurized pipe systems. This
hopefully leads to better design rules.
This thesis report is an investigation of one possible cause of air in sewer system
pipes, namely air intake at the pump inlet by air entrainment in the sewer sump. The
thesis is connected to the CAPWAT project, but is not a part of it.

Cause of the research


Air intake by sewer pumps is one of the main reasons that pressurized sewer systems
do not reach their design capacity discharges. Because of that, running and
maintenance costs are higher than necessary. But there is still little known about air
entrainment due to plunging jets from circular open channels, and air intake by
submersible pumps near plunging jets, especially at magnitudes of real sewer sumps.

Research objectives
The research objective is to achieve knowledge and insight into air entrainment from
free overfall water jets from horizontal open channels and air bubble intake with
submersible sewer pumps.

The main questions are:


• How does the air entrain into the reservoir?
• How many air bubbles can be expected where in a certain configuration
without any specific measures?
• What is the sphere of influence of the submersible sewer pump?
• What is a robust measure to prevent air coming into the sewer pump?

Report contents
At first a literature study about plunging jets and free overfall flows from circular
channels is given in Chapter (2). Chapter (3) describes the model set-up and the
measurement program and procedures. The qualitative findings about plunging jets
and bubble plumes are discussed in Chapter (4). The measured results are then
summed up in Chapter (5). The last two chapters contain the conclusions, Chapter (6),
and recommendations, Chapter (7).

10
Chapter 2: Literature study
After the introduction paragraph (2.1) a dimensional analysis is given in paragraph
(2.2). Then, paragraph (2.3) is about air entrainment mechanisms. The basic
hydrodynamic features of liquid jets and air bubbles are described in paragraph (2.4).
The following paragraphs give some entrainment features, like minimum entrainment
jet velocity (2.5), volumetric flow rate of entrained air (2.6) and air/water entrainment
ratio (2.7). The characteristics of bubble dispersion are summed up in paragraph (2.8).
Information about free overfall from circular open channels is available in paragraph
(2.9) and the last paragraph (2.10) gives the conclusions from the literature study.

2.1 Introduction
This chapter describes what is already known about air entrainment. First there is a
dimensional analysis about physical models and scaling in paragraph (2.2). Next there
will be some paragraphs about plunging jets from small nozzles (paragraph (2.3) and
(2.4)). After that the paragraphs (2.5) to (2.7) will tell something about air
entrainment, followed by the characteristics of bubble dispersion in paragraph (2.8)
and a last paragraph (2.9) about free overfall from circular open channels.

In the past, a lot of research has been performed on air entrainment with small
jets, produced by pressurized water supply systems with nozzles. In literature used
nozzle diameters have a range of 1 to 40 mm and the accompanying jet velocity range
is 1 to 37 m/s (Bin, 1993). This research project is different from all the projects
described in literature. De major difference is the use of a free falling jet, from a
horizontal open circular tube, instead of a water supply with a small nozzle.
Furthermore, the sewer pipe as used in practice has a diameter of several decimeters,
instead of the small nozzle diameters described above.
It is clearly that the information that has been written about fast and small liquid
jets in the literature, will give useful information about air entrainment, but most of
the times the empirical data do not apply for the dimensions used in this project.
Besides, some aspects typical for free falling jets from open supply systems are not
considered in literature, for instance the influence of the shape of the jet cross section
profile.

This chapter is mostly based on the review article of Andrzej K.Biń: ‘Gas entrainment
by plunging liquid jets’ (1993).

2.2 Dimensional analysis


When, in a research project, a physical model is used to measure the behavior of a
certain (future) real situation. It may be possible to make the model in another scale
than the real situation or prototype, especially when the scale of the prototype is very
large compared to the model test laboratory. When that is the case, scale effects will
possibly be apparent and has to be minimized as much as possible. These scale effects
will be small when the next dimensionless parameters are followed.

Following the Buckingham ∏-theorem (Chanson, 2004), there is a function with


five independent dimensionless parameters which decide if liquid flow scale models
have scale effects or not:

11
 
 2 
V ρV ρVL V V 
F ( Fr ; Eu; Re;We; Ma ) = F  ; ; ; ; =0
 gL ∆P µ σ Eb 
 
 ρL ρ 

In a geometrical similar model, true dynamic similarity is achieved if and only if each
dimensionless parameter has the same value in both model and prototype. When the
pressure difference ∆P is treated as a dependent parameter and the Sarrau-Mach
number Ma is omitted, because of the very small influence in both model and
prototype, the dynamic similarity in most hydraulic models is governed by:

 
 
V ρVL V  ∆P
F ; ; = ⇒ F ( Fr ; Re;We ) = Eu
 gL µ σ  ρV
2

 
 ρL 

The use of the same fluid on both the prototype and the model satisfies the Froude,
Reynolds and Weber number scaling criteria, because:

• the Froude number Fr similarity requires VR = LR


1
• the Reynolds number Re scaling implies that VR =
LR
1
• the Weber number We similarity requires: VR =
LR

Hereby is the scale multiplier LR and VR the ratio between model and prototype for
 l d H   V 
the length  LR = P = P = P  respectively velocity  VR = P  .
 lM d M H M   VM 
In most cases only the dominant mechanism is modeled:

• In fully enclosed flows (pipe flows), the pressure losses are basically related to
the Reynolds number Re, so a Reynolds number scaling is used (ReM = ReP).
• In free-surface flows, gravity effects are always important and a Froude
number modeling is used. Note that the model velocity is less than that in the
prototype for LR > 1 and the time scale equals: t R = LR . A Froude number
modeling is typically used when friction losses are small and the flow is
highly turbulent (ReM > 5000). Where the Reynolds number is defined in
ρVDH
terms of the hydraulic diameter ( Re = ). The model must have the same
µ
relative roughness ((ks)R = LR) as the prototype.
• In case of entrainment of air bubbles in free-surface flows, gravity effects are
predominant, but it is recognized that surface tension scale effects can take
place for LR > 10 (or LR < 0.1) or even less.

12
2.3 Air entrainment mechanisms
Several studies showed that air entrainment with plunging jets from nozzles, takes
place when the jet impact velocity exceeds a characteristic velocity. This velocity is a
function of the inflow conditions (McKeogh & Ervine, 1981). The mechanism of
bubble entrainment depends upon the following:
• Jet velocity at impact
• The physical properties of fluid (mainly viscosity)
• The jet nozzle design
• The length of the free-falling jet
• The jet turbulence (Bin, 1993).

2.3.1 Viscous and low velocity jets


For viscous and low velocity laminar jets, a collar like meniscus forms around the jet
at the plunging point. At slightly higher flow rates a depression in the pool surface
forms as a result of the impact pressure of the associated air boundary layer and from
the flow field induced in the pool itself. Where the jet meets the pool, it is enlarged to
a diameter 1.5-2.5 times the diameter of the approaching jet (see figure 2.1.a.). With
further increase in the flow rate, the depth of depression becomes greater and the
contact angle αj decreases towards 0º. The jet transfers its momentum to the pool,
creating transverse and normal velocity gradients near the surface. These give rise to a
viscous shear and normal stresses which tend to pull the liquid near the surface deeper
into the pool, resulting in an “inverted” meniscus around the junction. When the
contact angle reaches 0º, an air film is carried into the pool by the fast-moving liquid,
see figure 2.1.b. The lower part of this film oscillates and eventually breaks up into
bubbles. Now the velocity of the jet is just above that necessary for entrainment.

Figure 2.1: Viscous liquid jet plunging through a pool surface: (a) contact angle 0 < αj < 90º; (b)
formation of an air film (αj = 0º) (Biń, 1993)

Induction funnel
Entrainment from an induction funnel is more regular than under extreme conditions,
although still intermittent. A possible explanation for the change from a gas film
formation with low velocity jets to the induction funnel form is the “tearing of the
surface”. The tearing of the surface, which arises from the liquid entrainment in the
reservoir around the plunge point, is significantly assisted by the physical pressure
downwards on the meniscus as the contribution of the air boundary layer becomes
more significant. Near the limits of stability, small changes in the jet velocity can lead

13
to penetration and capture of air. It is evident that any large scale rotations in the
receiving pool will be amplified by the conservation of angular momentum in the
inflowing liquid near the pool surface. This will lead to free vortices at the plunge
point around smooth jets, which will deepen the induction funnel and makes the
contact angle smaller. This enlarges the chance to have a contact angle αj = 0º, and so
the chance to have air entrainment.
The intensity of the circulation leading to the formation of the induction funnel
will be strongly dependent on the geometry of the receiving bath, and in particular on
the presence of baffles or other obstacles that might hinder the development of large
scale rotations.

With larger and generally swifter jets, the more severe conditions lead to the
entrainment of considerable quantities of air. This gas rises immediately around the
plunge point, giving rise to a bubble column effect that is strong enough to produce a
reversal of the flow in the receiving pool; circulation at the surface is now outwards
from the jet towards the walls. Under these conditions it is impossible for the
induction funnel to form.

2.3.2 Jet surface instabilities


For a low viscosity jet, the mechanism of entrainment is related to the jet surface
instability and the interaction of disturbances on the jet surface with the receiving
pool liquid. The liquid surface has at first a small depression which is caused by the
impact pressure of the associated boundary layer gas and the flow field induced in the
pool liquid itself. When disturbances on the jet surface hit the depression, the surface
deformation, transverse bulk liquid movement and the resulting formation and closure
lead to gas entrainment at the plunging point, see figure 2.2.

Figure 2.2: Air entrainment mechanism: (a)-(d) show subsequent phases of the phenomenon as a
disturbance in the jet moves downwards (Biń, 1993)

The horizontal movement on the free surface is not fast enough to follow the
roughness of the jet as it passes by, resulting in the capture of gas bubbles.
In real situations the successive disturbances on the jet produce an irregular gas
entrainment. With longer jets the surface irregularities become bigger, producing
greater entrainment.

14
2.3.3 Other entrainment mechanisms
Disrupted jet
Long jets eventually lose coherence and high velocity jets frequently experience
partial disruption as a result of the air friction forces. Individual droplets impinging on
the liquid surface will entrain air if their velocity is high enough (above about 1 m/s).

Entrainment with surface foam


There is another mechanism that has been suggested by Bin (1993): Most of the
entrained air enters the main flow via the layer of foam forming on the surface of the
receiving fluid. Maybe that is because of the local amounts of surface active agents,
which decreases the surface tension of the water and result in smaller entrained air
bubbles.

2.3.4 New findings on entrainment mechanisms


A relative new experimental research by Zhu et al.(2000) shows that, provided the
nozzle is properly contoured, an undisturbed jet does not entrain air, even at relative
high Reynolds numbers. They found that a plunging jet does only entrain air when the
jet has surface disturbances. Zhu claims that earlier conclusions were suggested by
experimentations with nozzles that, unlike Zhu’s, had not been designed to minimize
jet turbulence.

2.4 Basic Hydrodynamic features of liquid jets and bubbles


Research with plunging jets has, in the main, produced a series of empirical
relationships containing the major parameters, usually in the form of dimensionless
numbers. In this section, the basic hydrodynamic features about plunging jets will be
explained.

2.4.1 Impact velocity of the jet


1
For simplicity, the jet velocity from nozzles can be calculated as: V j = (VN2 + 2 gL j ) 2 .
For inclined jets the jet velocity at the plunging point can be calculated with the
equation above substituting for Lj the height of the nozzle outlet above the pool liquid
level.

Figure 2.3: Schematic representation of a jet from a nozzle (Yamagiwa et al., 1993)

15
According to Nakasone (1987) plunging nappes have a impact velocity of;
V0 = 2 g ( H j + 1.5hc ) , with Hj as the fall height and hc as the critical depth before the
falling edge.

2.4.2 Water jet diameter


A main problem is the calculation of de diameter and the velocity of the jet at the
plunging point, because gravity affects these parameters along the jet.

Laminar vertical jets


For laminar vertical jets, an energy balance, neglecting the effects of the surrounding
gas, leads to the equation for diameter (by intermittent velocities) (Bin, 1993):
1

dj  π 2 gL j d r4  4
= + 1 where dr is the diameter at Lj = 0.03m. The jets velocity profile
d r  8Qw2 

relaxation does not extend far beyond Lj/dN = 3.

High velocity jets


High velocity jets are subject to violent gas friction. This leads to surface tension
forces no longer dominantly controlling the surface character, so the jet surface
becomes ill-defined. Long exposure time photographs show that the jet spreads in a
dispersed, generally conical form.

Van de Sande and Smith (1973) suggested the following function:


dj m
= C (Wea Relength ) . With C = 0.085 and m = 1/6. This function is only valid for
dN
Wea * Relength > 7 * 105 and with long nozzles (lN/dN ≥ 50). Relength defines the
VN L j
development of the accompanying air boundary layer; Relength = and Weber
νw
VN2 d j ρ a
number; Wea = .
σ
According to Biń (1993) it is likely that the factors C and m depend on the nozzle
design or the developed turbulence level in the jet.

2.4.3 Jet surface roughness


It is likely that the nozzle design has an effect on the jet surface roughness. McKeogh
and Ervine (1981) studied the surface roughness using nozzles with different inner
diameters (from 6 to 25 mm), jets with different initial velocity (from 2 to 5 m/s) and
different levels of turbulence. Later Ohkawa et al. (1986) described this correlation in
a more feasible form:

0.9
2ε  Lj 
= 5.98*10−3  
dN  dN 

16
Lj
This function is valid for a ratio from 6 to 100. Where ε is the surface
dN
disturbance on the jet, see figure 2.4.

Figure 2.4: Surface disturbance on jet surface (McKeoch & Ervine, 1981)

It is evident that the jet geometry will depend on the turbulence level in the jet
itself, which is for its part dependent upon the nozzle design and all other factors
responsible for providing turbulence in the jet.

2.4.4 Break-up length of liquid jets


A possible mechanism of gas entrainment is when individual droplets impinge on the
liquid surface. These can also result from the break-up of the jet. Therefore, the
problems inherent to jet stability are relevant for gas entrainment by plunging liquid
jets.

For the laminar jet region, Bin (1993) suggested a relationship for the break-up
length:

LB 0.85
= 19.5WeN0.425 (1 + 3Z )
dN
Which is valid for WeN (1 + 3Z ) = 3 − 100 with Z as the Ohnesorge number
 µw 
=  , which relates the viscous and surface tension force.
 d N ρ wσ 
 

LB
For the turbulent jet region the relationship = CWeNp is most frequently
dN
proposed, with p equal to 0.31-0.32 for jets issuing from nozzles with lN/dN ≥ 5.
McKeogh and Ervine (1981) correlated LB with the discharge rate of water;
LB = CQwx . The power exponent x was found to be dependent on the turbulence level
in the jet and independent of de nozzle diameter.

17
2.4.5 Boundary layer of a surrounding gas
As the liquid jet moves through a gaseous atmosphere, a boundary layer will develop
along the jet, similar to that observed for a moving continuous cylinder with diameter
equal to dj and velocity equal to VN. Several complex approximated models relevant
to this situation are available, see Bin (1993).

2.4.6 Bubble rise velocity


Small air bubbles (da < 1 mm) act as rigid spheres. The motion of these small air
bubbles is dominated by the balance between the viscous drag force and the buoyant
force. For very small bubbles (da < 0.1 mm) the bubble rise velocity ur is given by the
Stokes’ law (Chanson, 1994):
2 g ( ρw − ρa ) 2
ur = da
9 µw
For small rigid spherical bubbles (0.1 < da < 1 mm), the rise velocity is best fitted by:
g ρw 2
ur = da
18µ w
As far as the fluid viscosity is neglected (da > 1 mm) the bubble rise velocity can be
estimated as:
2.14σ
ur = 0.52 gd a
ρwda

2.5 Minimum entrainment velocity


Bin (1993) describes that a simple correlation of all the available experimental data
between the minimum entrainment velocity Vmin and the Lj/dN ratio, is not possible,
see figure 2.5 for a chart with experimental results of many researches.

Figure 2.5: Impression of the large scatter of the experimental data points between the minimum
entrainment velocity Vmin and the ratio Lj/dN for the continuous jet region (Bin, 1993)

18
The values of Vmin, obtained at different turbulence intensities, are indicated for
comparison. It can be concluded from this figure that a large scatter of the
experimental data points is evident for jets produce from long cylindrical nozzles. The
length of the cylindrical section of the issuing nozzle is responsible for the level of
turbulence produced in the jet. The effect of the liquid viscosity seems to suppress the
turbulence.
0.164
 Lj 
For large nozzles (dN ≥ 7 mm) an empirical correlation; Vmin = 1.4   can
 dN 
be recommended, which seems to be applicable for short cylindrical nozzles (lN/dN ≤
3) and within Lj/dN = 1-100. For such nozzles the jets produced from them are of
V′ 
relatively low turbulence levels  j < 3%  .
 Vj 
 

Maximum impact angle with air entrainment


Experimental results show that more air will entrain with inclined jets. Detch and
Sharma (1990) studied the maximum angle of impact which would result in bubble
appearance for a given jet velocity. The experiments performed by these authors
covered nozzles with diameters from 1 to 4.6 mm and different liquids. Detch and
Sharma (1990) obtained a dimensional relationship between the critical angle and the
jet parameters (Vj, µw, ρw, σ):
 σVj 
α c = −242.13log10   − 79.1
 µw ρw 
σVj
The apparent range of validity of the equation is for values of the term from
µw ρ w
5 2
0.01 - 0.5 m /kg/s , Reynolds numbers between 800-10.000 and αc ≤ 60°.

2.6 Volumetric flow rate of entrained air


2.6.1 Methods of entrained gas measurements
Since gas entrainment by plunging liquid jets occurs as a localized phenomenon at the
plunge point, two groups of methods of entrained gas flow rate measurements were
developed:
• Catching gas after it has been entrained into the pool liquid (bubble trap)
• Measuring the removal of gas from a gaseous space above the pool surface
around the plunge point.

A bubble trap is a device what collects and measures the amount of air that
withdraws from de reservoir water, after be entrained by an oblique jet, see figure 2.6.

19
Figure 2.6: Schematic representation of a bubble trap (McKeoch & Ervine, 1981)

See figure 2.7 for a gas removal measuring apparatus.

Figure 2.7: gas removal measuring apparatus (Kusabiraki et al., 1990)

Each of the measuring techniques has its own shortcomings. In the case of
inclined jets, one usually has to carry out experiments with different angles of
inclination of the jet and extrapolate the data to the vertical position. Both traps and
gas removal arrangements may interfere with the fluid flow in the pool.

2.6.2 Air entrainment regions


Different studies are carried out by Van de Sande and Smith (1976). They describe
three regions in the entrainment curve defined by jet plunge velocities:
• Initial and/or low jet velocity region
• Transition region
• High jet velocity region

Thus a typical S-shaped entrainment rate curve is observed for the whole range of
jet velocities, see also figure 2.7.

20
Figure 2.7: Air volume entrained as a function of jet diameter and velocity (Sande & Smith,
1973)

Low velocity jet region


Van de Sande and Smith (1976) stated that the low jet velocity region extends up to
about 5 m/s. Others agree with this limiting value of Vj with calculations of empirical
expressions.
The plunge velocity of jets from horizontal channels depends mostly on de fall
height and the fall distances H < 1.0 m. For simplification reasons, this research
project will operate in this low velocity region.

The rate of gas entrainment by plunging liquid jets of the length near the break-up
length (≥ 0.9 LB), is directly related to the kinetic energy of the jet (Sande & Smith,
1976):
1 π 
Qa = function  ρ w d 2j V j3 
2 4 
See figure 2.8. In this case the jet behaves like a train of drops.

21
Figure 2.8: Amount of entrained air near the break point with αj = 60º jets (Sande & Smith,
1976)

For jets with lengths less than 90% of the break-up length, Sande & Smith (1976)
suggested an air entrainment rate:
Qa = 0.015 X 0.75
−1.5
With X = d N2 V j3 L0.5
j (sin α j ) and X = 10−4 − 10−2 . The experiments were carried out
with jets produced from nozzles with dN = 2.85 – 10 mm, lN/dN = 50, Vj = 2 – 5 m/s,
Lj < 0.5 m and αj = 20 - 60º.

High velocity jet region


The high jet velocity region was fixed by Van der Sande and Smith (1973) at the
value of Wea > 10. The more recent data of Kusabiraki et al (1990), show that the
transition to this region is dependent on the experimental conditions (dN, Lj, lN/dN, αj,
νw).

High velocity jets not only carry air captured within the mean containing
envelope, but also entrain air in the boundary layer that develops outside that
envelope into the pool. Thus the total amount of entrained air is made up of two parts
(Bin, 1993):
Qa = Qa1 + Qa2.
πV
The portion of air captured by the jet roughness is given by: Qa1 = N ( d 2j − d N2 )
4

whereas the boundary layer can be expressed by: Qa 2 = ∫ V 2π rdr , whereby Va is
dj 2
a

the local velocity of the boundary layer air at a given radius measured from the jet
axis.

22
It should be emphasized that the contribution of the boundary layer to the total
amount of entrained air depends upon the situation. Normally, it is between 20 – 70%
of the total and therefore cannot be neglected.

Transitional jet region


The transition from the low velocity region to the high velocity region is continuous.
Bin (1993) describes a simple procedure to predict the entrainment rate for this
region, based on the interpolation between the amount of entrained air at Vj = 5 m/s
and from the jet velocity corresponding to Wea = 10. The error with this procedure is
within ± 20%.

2.6.3 Effect angle of impact on air entrainment


An obliquely impacting jet introduces a horizontal velocity component in the liquid
and the hole created by the jet in the liquid gets larger. The vacuum increases and
more air will entrain. Experimental results confirm this hypothesis. Measurements
done by Sande & Smith (1973), at impact angles from 60, 70 and 75º show no further
significant angle dependency and may also be taken as representative for vertical jets.
That is because measurements of vertical jets cannot be made satisfactorily with a
bubble trap apparatus.
In spite of a slight dependency of Lj, Vj, d j and probably the characteristics of the
1
fluid, a sufficiently accurate relation for the angle dependency is; Qa = C (sin α )− 4 ,
with C a constant.

2.7 Entrainment ratio Qa/Qw


Entrainment ratio is defined as the ratio of the volumetric flow rates of air and water
and is a measure of performance of the plunging jets. According to measurements of
many authors, the entrainment ratio will depend on the basic system parameters, such
as jet velocity, jet length, nozzle diameter and design, angle of jet inclination and the
physical properties of the water. In order to quantitatively simplify the effect of
different parameters on the entrainment ratio, a correlation between this ratio, the jet
Froude number and the ratio Lj/dN has been sought.

2.7.1 Entrainment ratio Qa/Qw for different jet regions


An empirical expression for high velocity liquid jets plunging though a gaseous
atmosphere into a pool is the following ratio for gas flow to liquid flow (Burgess,
1972):
2
Qa  d j 
=  −1 ,
Qw  d N 
Where dj and dN are the jet diameter at plunge point and nozzle respectively, see also
figure 2.9.

23
Figure 2.9: Specification of jet surface roughness (Burgess et al., 1972)

For vertical jets, Bin (1993) proposed the following empirical expression for all
regions of entrainment:
Qa 0.4
= 0.04 Frj0.28 ( L j / d N ) .
Qw
It corresponds satisfactory with experiments of other authors provided that Lj/dN ≤
0.4
100, lN/dN ≥ 10 and Frj0.28 ( L j / d N ) ≥ 10 .

A re-analysis of previous work by Chanson (1994), shows that the dimensionless


quantity of air entrainment can be estimated as:
Qa Qa Fr 2
= C * Fr 2 or =C* 1.2
for Vj < 5 m/s
Qw Qw ( sin α )
Qa 1
=C* for 5 < Vj < 10 m/s
Qw Fr
Qa
= C * Fr for Vj > 10 m/s
Qw
V − V0
Where Fr is the jet Froude number defined as Fr = and V0 is the velocity at
gd j
which air entrainment commences. V0 is almost constant for large turbulent intensities
(Tu > 3%), with typical values ranging from 0.8 to 1 m/s.

24
2.7.2 Effect of surface tension and liquid viscosity on entrainment
ratio Qg/Ql
Several authors studied the influence of liquid viscosity and surface tension on the air
entrainment rate. In these studies the liquid phase kinematic viscosity νw was varied
from 8.8*10-7 – 1.5*10-5 m2/s, whereas the surface tension σ varied from 0.024 -
0.076 N/m.

Above about µw = 6.5 mPa s, the shear between the pool liquid and the jet,
controls the entrainment, whilst below this value, the roughness of the jet and the air
boundary layer are the controlling parameters.

Bin (1993) presents the experimental data of Kusabiraki et al. on Qg/Ql with
respect to a term which combines the physical properties of the liquid phase
 µl 
  . Both researchers discovered three regions of such dependence at the same
 ρσ 
 l 
values of dN, VN, Lj/dN, and lN/dN, see figure 2.10. The transition values of µl / ρlσ
between the regions were 1.2*10-4 and 1.9*10-4 m1/2.

Figure 2.10: Dependency of Qg/Ql and µ l / ρl σ (Yamagiwa et al., 1993)

The Qg/Ql values were correlated by (Kusabiraki, 1990) in the following form
within a 20% error.
B
(L / d N ) ( lN / d N ) Z D ( sin α )
C E
G
Qg / Ql = AFr F (sin α ) j

The empirical constants A – G, which do change according to the range of lN/dN and
VN, are presented in (Yamagiwa et al., 1993)

25
2.8 Characteristics of bubble dispersion
2.8.1 Bubble size distribution
As a result of gas entrainment by plunging liquid jets, bubbles are dispersed below the
pool liquid surface. The dispersed bubbles form two distinctly different regions, see
figure 2.11:
• A characteristic biphasic conical region comprising fine bubbles with
diameters less than 1 mm.
• A region of bigger rising bubbles (secondary bubbles), which surrounds the
former one.

Figure 2.11: Schematic representation of biphasic conical region and the rising bubble region
(McKeoch & Ervine, 1981)

The structure of the biphasic cone and the surrounding bubble column is complex.
In the cone, high turbulence intensities and shear stresses, created by the bubbles,
break the captured gas into fine bubbles. Bubbles escape from the cone at its
boundaries and at the bottom as the buoyancy forces overcome the jet momentum and
the liquid local velocity decreases. Coalescence also takes place in these sections of
the cone. The largest recorded bubbles in the water had diameters of about 7 mm.

There is general agreement that for the air-water system, bubbles formed during
entrainment by vertical or inclined jets in the rising bubbles region (secondary
bubbles) have (Sauter) diameters of about 3-4 mm, practically independent of the
experimental conditions. The bubble size variation distribution is approximately
normal. In the biphasic cone, bubbles are much smaller than the bubbles in the rising
bubble region (< 1.4 mm).

Bin (1993) describes a quasi-theoretical relationship between the volume-


equivalent bubble diameter and the entrainment ratio for the air-water system:
d v = 4.3*10−3 3 Qa / Qw .

26
2.8.2 Penetration depth of entrained bubbles
Bubbles entrained by a vertical plunging jet penetrate the pool liquid to a maximum
depth. This point is not strictly defined since the lower limit of the bubble swarm
fluctuates continuously, but a time average can be estimated. Several authors
measured the maximum depth of bubble penetration in the vertical or inclined
plunging jet systems. At the maximum depth of bubble penetration, the local liquid
velocity in the submerged jet at that point is assumed to be equal to the bubble free
rise velocity. This led to a direct linear relationship between the maximum penetration
depth and the product of the jet diameter and jet velocity.
Biń (1993) mentions a simple pure empirical relationship for the maximum
penetration depth:
H p = CVNn d Np .

If all available experimental data on Hp are considered, then for VN dN ≥ 0.01 m2/s, n
= p = 0.66 and C = 2.4, whereas for VN dN < 0.01 m2/s, n = p = 1.36 and C = 2.4.

Buoyancy forces of air bubbles


Following Sande & Smith (1975), this simple momentum approach is insufficient,
because the amount of entrained air can be different by, for example, other jet lengths.
With more entrained bubbles, the total resistance and buoyancy forces of the bubbles
are no longer the same at the impact point. By considering the shape of the bubble
cone and the depth of the penetration, it is observable that the cone has become wider
and the depth has decreased. They suggested the expression for vertical jets:
4
Vj 3d j
H p = 0.42 , which has a 10% accuracy with their experimental data, obtained
4 Q
a

with nozzles of diameter ranging from 3.9 to 12 mm and for Lj < 0.5 m.
Under practical conditions Hp will not be greater than 0.4-1.0 m.

Nakasone (1987) suggest that the penetration depth of entrained air produced with
nappe jets is about: Hp = ⅔Hj. See also figure 2.12.

Figure 2.12: Schematic drawing of the penetration depth of a nappe (Nakasone, 1987).

Following Yamagiwa (1993) the dimensionless maximum bubble penetration


depth Hp/dN is hardly affected by liquid properties. For liquids in all three regions, see
paragraph 2.7.2, the following correlations were found:

27
a
H p /d N = ( L j /d N ) ( LN / d N ) 10 f ( Fr ) (sin α )c
b

2
f ( Fr ) = d + e ( log Fr ) + f ( log Fr )
The empirical constants a – f are presented by Yamagiwa (1993).

Effect of nozzle design


Ito et al. (2000) give a similar empirical formula for the penetration depth with other
constants than Yamagiwa (1993). They also take into account the nozzle contraction
angle between the feed and the nozzle diameter, see figure 2.13.

Figure 2.13: Schematic drawing of nozzle design (Ito, 2000)

Maximum bubble penetration depth with jets from nozzles


A theoretical value of the maximum bubble penetration depth can be deduced from
the continuity and momentum equations for diffusing jets. Assuming that the bubbles
are entrained to a depth where the vertical component of the mean jet velocity equals
the bubble rise velocity, Chanson (1994) gives a method for inclined plane and
circular jets. For plane jets, the penetration depth Hp is correlated by:
2
2  2 
Hp  V j  sin 3 θ   ur  tan β 
= 0.0240   1 + 1 − 20.81  , where β is the spread angle
 V  sin 2 θ 
 ur  tan β 
2
dj  j  
 
of the bubble cone in the fully developed flow region. For circular jets, β can be
estimated around 14 degrees on both models and prototypes.

2.8.3 Aeration and bubble zone


The bubble zone is the horizontal distance from the edge of the weir or tube where the
jet starts to fall, to the distance where no air is in the water, see figure 2.13 or
appendix A. The aeration zone La is the bubble zone Lb minus the horizontal fall
distance Ljx.

Figure 2.13: Schematic representation of bubble respectively aeration zone at deep or shallow
reservoirs (Nakasone, 1987)

28
The length of the bubble zone, by means of photographs at many different
experiments with water nappes, the following empirical formula for the length Lb of
the bubble zone was obtained by Nakasone (1987).
0.134
Lb = 0.0629 ( H j + 1.5hc ) qw0.666
Hereby is qw the discharge per meter width of weir in m3/s/m. (In the paper of
Nakasone (1987) on page 74, the dimension of q is written as m3/h/m, but after
verification with values from this experiment and checking his list of symbols, this
should be m3/s/m) This formula is determined with a shallow pool of 0.50 m deep.
For larger reservoir depths whereby bubbles do not reach the bottom, the length of the
bubble zone will probably be shorter.

2.9 Free overfall with circular open channels


2.9.1 Introduction
Figure 2.14 present a schematic view of a typical free overfall including the hydraulic
aspects and the variation of streamlines curvature of an overfall.

Figure 2.14: (a) Schematic view of a typical free overfall and the hydraulic aspects; and (b)
streamline pattern of a free overfall (Dey, 2002)

The flow in the vicinity of the channel edge is affected by the vertical acceleration
due to a non-hydrostatic pressure in the water column, see figure 2.14.a. This results
in an end depth being less than the critical depth. In mildly sloping channels, the
approaching flow is sub-critical becoming supercritical just upstream of the channel
end section.

For all overflow geometries, there is a unique relationship between the end depth
and the critical depth and between the critical depth and the water discharge. Only
when the approaching flow is supercritical, the critical section does not exist upstream
of the channel end section. The discharge is than a function of the end depth and the
channel slope.

29
2.9.2 End Depth Ratio he/hc
Dey (1998, 2001 and 2002) presents a simplified approach to determine the end depth
of a free overfall in horizontal or mildly sloping circular channels. Following Dey
h h
(2001) the end depth ratio EDR is between 0.72 < e < 0.74 for c < 0.86 . An often
hc D
used EDR from Rajaratnam & Muralidhar (1964) is equal to 0.725. Aigner uses a
ratio of 0.742.

2.9.3 End Depth Ratio with sloping channels


Rajaratnam and Muralidhar (1964) experimented also with sloping channels. The
results are shown in figure 2.15. The critical slope Sc for the pipe, was calculated with
h S
Manning’s n = 0.013. Notice that e has a systematic variation with .
hc Sc

Figure 2.15: End Depth Ratio for circular channels (Rajaratnam and Muralidhar, 1964)

In this figure can be seen that a little positive inclination has more effect on the
end depth ratio than a little negative inclination. This counts too for the approach flow
velocity.

2.9.4 Dimensionless discharge


Aigner has formulated a semi theoretical formula for the dimensionless discharge:
11 11
Qˆ w = 1.371hˆe6 = 0.794hˆc6
Qw h
With dimensionless parameters: Qˆ w = and hˆ = . Hereby is D the diameter
2.5
gD D
of the tube. The formula has an error of 5% with empirical solutions of other authors
who found formulas with other constants a and b in: Qˆ w = ahˆeb .
Rajaratnam & Muralidhar (1964) supposes the following constants: a = 1.54 and
b = 1.84 ≈ 11/6.

30
2.9.5 Properties of circular open-channels
This section contains a list of circular open-channel properties, presented by Chanson
(2004). Here D is the tube diameter and δ the angle from pipe centre to the
intersections between water surface and tube wall (can be >180 degrees), see also
figure 2.16.

Figure 2.16: Definition sketch of a circular channel cross section

D δ
• Flow depth: h = 1 − cos 
2 2
 2h 
• Angle: δ = 2 cos −1 1 − 
 D
δ
• Free-surface width: B = sin
2
• Cross-sectional area:
D2 D2  h 
(δ − sin δ ) =  arccos 1 − 2  − 2 1 − 2  1 −  
h h h
A=
8 4   D  D D D
D
• Wetted perimeter: Pw = δ
2
 sin δ 
• Hydraulic diameter (equivalent pipe diameter): DH = 4 RH = D 1 −
 δ 
VDH
• Reynolds number for pipe and open channel flows: Re =
ν

31
2.10 Conclusions literature study
The many mechanisms of gas entrainment by plunging liquid jets make it hard to give
a quantitative prediction of the performance of the plunging jet system, at least in
terms of the primary variables (jet diameter, jet velocity and length, as well as the
physical properties of the fluid). Many secondary factors (nozzle design, angle of jet
inclination, presence of vibrations) can have significant influence on jet behavior, but
these are even more difficult to quantify.

Many of the mentioned functions and relations are based on experimental data for
fast and small jets from nozzles. Therefore all except the circular channel formulas are
only applicable for a scale order much smaller than the scales used in real sewer
systems. The objective of this thesis is to see if the relations described in literature
may be used for free overfall jets from circular channels with much bigger
dimensions. If that is not the case, some minor changes may be enough to still use the
functions.

In addition to the verification of the formulas described in literature, this report


will give a qualitative description of a free overfall jet from a circular channel,
because there is no information available about this specific subject.

The objective of this report is to find relations between plunging jets, air bubble
entrainment and the air bubble intake with a submerged sewer pump in a small sewer
sump. There is no information available for this subject too, which means there is still
a lot to discover about this subject.

32
Chapter 3: Description of experimental model set-up
and measurement program
The first paragraph (3.1) tells something about the used model reservoir. Then the
horizontal circular approach channel is described in paragraph (3.2). The submersible
sewer pump and the flow regulation follow in paragraphs (3.3) respectively (3.4).
Paragraph (3.5) is about the measurement attributes and the last paragraph (3.6)
describes the measurement program, procedures and parameters.

3.1 Model reservoir


After examining examples of sewer sumps in practice, the set-up of the model is
determined. It appears that sewer sumps exist in many different shapes and
dimensions. The amounts of sewer pumps vary too. The research will be more
fundamental when the sump is block shaped and contains only a single pump.

Figure 3.1: Schematic drawing of the model set-up (See appendix A for an enlarged version)

Definitions in figure 3.1:


• Aeration zone La (m)
• Bubble zone Lb (m)
• Fall distance Ljx (m)
• Tube–pump distance Ltp (m)
• Critical depth hc (m)
• End depth he (m)
• Fall height H (m)
• Real fall height He (= H + 2/3 he) (m)
• Penetration depth Hp (m)
• Water depth Hw (m)
• Total plunge velocity Vj (m/s)

33
An approach flow tube in the model, see figure 3.1, represents the mouth of a
(small) sewer system that ends in a sewer sump. In normal dry wetter conditions, the
mouth of the sewer system is always above the water level in the reservoir. This way,
sewer systems have always their maximum buffer capacity in case of a short but
heavily rainfall. Secondly, the low water levels in sewer pipes will wash away
sediments in the sewage and will not be able to settle and block the sewer pipe, as
flow velocities are high at low water levels in the pipe.

The tube and the pump are lined up in the middle of the reservoir to get a
symmetrical reservoir. To keep the model manageable, the set-up must have the
dimensions of a relative small sewer sump, and if possible without any scaling, to
exclude any possible scale effects. More about this subject further on in this
paragraph.

TOP sewer sumps


The internationally acting concern ITT Flygt has developed a new prefab sewer sump
called TOP, see figure 3.2. The relative small sewer sumps are available in different
volume sizes; from ¾ to 15 m3 and have diameters from 800 to 1800 mm (Flygt).
These sumps act as good examples for the sewer sump model.

Figure 3.2: Schematic representation of an ITT Flygt TOP sewer sump.

Except for the smallest, all TOP sewer sumps contain two pumps to pump the
sewage to a cleaning station. The reason for that is the possible failure of one of the
pumps in case of blockage with big or tough objects, but the main reason is to
minimize the amount of pump restarts. The amount of restarts per hour is limited, due
to technical reasons. But still the model set-up has only one single pump, for
simplicity reasons.

34
Dimensions of sewer sump model set-up
For experimental researches with air bubbles, it is nearly impossible to obtain real and
valid results with scaled models. The viscosity and the surface tension of the water are
not as easy scalable as model lengths and velocities. For that, it is better to exclude
any scale effects by employ no scaling between prototype and model. For more
information about scaling, see paragraph 2.2 for a dimension analysis.
Without scaling, sewer pipe diameters can be large in practice. But to keep the
model manageable, a small but real sewer pipe diameter of 0.20 meter is chosen as
approach flow channel.

A part of a large flume of WL|Delft Hydraulics is used as a reservoir for the


model. The flume is made of glass for a good sight on the occurring phenomena. The
section dimensions of the flume are: h * b = 1.0m * 1.0m and it has a length of 110
meters. In this way, the length of the model reservoir can be adjusted easily and
almost infinitely by movable sheets. The installations to fill and empty the flume are
already present, so the ability to use this flume is a perfect opportunity for this
experimental research project.

3.2 Horizontal circular open tube as approach flow channel


The approach flow channel for the free overfall jet is a 5.0 meter long horizontal PVC
tube with (outer) diameter of 0.20 meter. This tube represents the mouth of a small
sewer system that ends in a sewer sump.

Adjustable approach tube height


According to the rule of thumb from Nakasone (1987), the penetration depth is about
2/3rd of the falling height, see paragraph 2.8.2, and the reservoir depth is 1.0 meter.
This leads to a maximum fall height of 1.0 meter above maximum water level. The
approach flow tube must then be able to set on 2.0 meters above the reservoir bottom.
The minimal fall height of the jet is 0.40 meter above the reservoir bottom.

Because the supporting structure of the approach tube is constructed on the rails
of the flume, it is possible to vary the distance between the pump and the tube mouth.
The length of the reservoir itself is also adaptable by moving a non-waterproof
plywood sheet in the flume. The reservoir walls do not have to resist high forces, so
this is the most easily and quickly way to displace it.

3.3 Submersible sewer pump


To appoint a pump type with the sufficient capacity for this research, the following
calculation is done to determine the maximum design discharge.
Because there is not often a super critical flow in pipe sections in sewer systems, the
formula V = gr is used as the maximum flow velocity in open circular channels,
with ‘r’ the radius of the tube. The maximum velocity is then:
Vmax = gr ≈ 10*0.1 = 1.0 m/s. The flow cross section area is:
A = π r 2 = π *0.12 = 0.031 m2, so the maximum discharge in a circular open tube is
about: Qw,max = Vmax A ≈ 1*0.031 = 0.031 m3/s = 31 l/s.

35
By knowing this maximum design discharge, ITT Flygt offered a submersible
sewer pump with a capacity of 30 l/s and a water head of 6.0 meters, which is
sufficient for the model set-up.

3.4 Types of flow regulation


Flow regulation by a control valve
The water that is taken in by the pump flows through a pipe system to the approach
flow tube and plunges back into the reservoir, see figure 3.1. Before the water is
transported to the approach channel, it passes a flow meter and a valve to measure the
discharge. Because the valve is computer controlled, it is possible to keep the water
discharge to a constant value, but it is possible to switch off the automatic valve
control.

With this model set-up it is only possible to measure with the same in- and
outgoing discharges, which is almost never the case in practice. In practice the
outgoing pump discharge is much bigger than the incoming flow into the sewer sump
at dry weather. The pump impeller frequency and pipe characteristic are always
constant and designed to handle rain conditions. When the water level reaches the
switch off threshold, the pumps are shut down. This is a difference between the model
set-up and reality.

Flow regulation by pump frequency transformer


In the model set-up, the pump is connected with a frequency transformer. This
transformer is able to change the rotation frequency of the pump impeller. Thus with
the same pipe system and valve position, the flow in de model is adjustable. In real
situations, a sewer pump can only pump at its maximum frequency, but this
transformer gives the possibility to change independently parameters.

3.5 Specific measurement attributes


Attributes for pump air intake discharge measurements
To measure the total volume of the air bubbles that is taken in by the pump, it is
necessary to collect all the air bubbles in the pipe system. The best and easiest way to
do this is with a closed vertical stand pipe connected to the upper side of the
pressurized pipe, see figures 3.3 and 3.4. This method only works under two
conditions: First, all air bubbles must be able to rise to the upper side of the horizontal
pipe before they pass the position of the vertical stand pipe and secondly the flow
velocity in the horizontal pipe must be low enough to catch all air bubbles on the
upper side of the pipe with the stand pipe. An estimation for the maximum pipe water
velocity is 10 D per second, with D the stand pipe diameter.

The volume in the vertical Perspex stand pipe is known by measuring the height
of the air volume in de stand pipe. Along with the measured pressure in the air
volume, is it possible to calculate the volume at atmospheric pressure.

36
Figure 3.3: Vertical stand pipe made of Perspex which takes air bubbles out of the pipeline, see
air bubble at red arrow

In this set-up it was possible to catch all air bubbles with a 5 meter long horizontal
pipe, see figure 3.4, with a diameter D = 0.20 m and a discharges Qw ≤ 12.5 l/s. This
way, the smallest air bubbles needed a rise time of 12.5 s to overcome D = 0.20 m,
which is equal to 1.6*10-2 m/s.
Air bubbles rise quicker when there are no violent turbulent eddies in the
horizontal pipe. This is because these eddies are able to mix especially the smallest
bubbles in the pipe. Therefore a sheet with small holes was placed at the front side of
the horizontal pipe section to suppress big eddies in the turbulent pipe flow.

37
Figure 3.4: Horizontal pipe section to let raise all the air bubbles to the upper side before
collecting the bubbles in the Perspex vertical stand pipe on the right.

Entrained air volume measurement


There are two known methods to measure the amount of entrained air with plunging
jets. The first one is collecting and measuring the air bubbles that withdraw from the
reservoir, see figure 3.5.a. All the bubbles are collected by a bubble trap above the
water and the air volume is measured by an air flow meter.
The second method measures the amount of air that is entrained and withdrawn from
a space around the plunge point, see figure 3.5.b. Method (a) is chosen for this thesis
because of the possible influence of the space holder around the plunge point on the
air entrainment with method (b).

The walls of the bubble trap used are 0.15 m below the water surface. So the
maximum pressure below the bubble trap is the pressure of a water column with a
diameter of 0.15 m, which is about 1.5 kPa. Through the narrow passage in the used
air flow meter, the maximum possible air discharge is equal to Qa = 100 l/min.
Unfortunately, this was not the maximum air entrainment that occurred with this
model set-up, so not all entrainment values could be measured.

38
(a)

(b)
Figure 3.5: Schematic representations of possible air entrainment measurement techniques (a):
(McKeogh and Ervine, 1981), (b): (Kusabiraki et al., 1990)

3.6 Mesurement program and procedures


This paragraph describes first the measurement program (3.6.1) and the measurement
procedures (3.6.2). In paragraph (3.6.3) is a list of all imposed parameters and the
measured parameters are summed up in paragraph (3.6.4).

3.6.1 Measurement program


The first experiments are done with a deep reservoir (0.95 m) and with no effect of
the pump or reservoir wall at respectively down- and upstream end of the reservoir.
So only the bottom and the side walls of the sump have any but minimal effect on the
air bubble plume.

The reservoir depth is decreased to see what the effects are on the air bubble
plume. In these experiments only the bubble zone and the penetration depth are
measured, which was most times to the bottom with the small water depth of 1.0
meter.

Then the air entrainment measurements are done with the bubble trap. So the
effects of the different discharges and falling heights are measured on the total
amount of air entrainment.

The effect of the pump impeller revolution rate alone is measured. It became
apparent that it was certainly important, which was unexpected.
In the last experiments the pump air intake discharges with respect to different water
discharges, fall heights, tube – pump distances and water depths are measured.

39
All experiments are measured only in stationary situations with stationary
parameters, like: water level, water discharge, etc. Real situations like plunging jets
with dropping reservoir water levels, are not taken into account, but are nevertheless
very interesting.

3.6.2 Measurement procedures


The measurement procedure was as follows: a water level, fall height and discharge
are set and the automatic control of the valve is switched off, so the valve position is
constant during the measurements. That is especially important when a lot of air
bubbles will be taken in at the pump and the discharge drops as well as the pump
efficiency.
With this set-up the penetration depth Hp, fall distance Ljx and the bubble length
Lb are measured on sight with a block against the glass wall to see if the line of sight
is perpendicular to the flume wall. The distances are then determined with a ruler
fixed to the glass wall. The measured lengths of the air bubble plume are based on the
rate of exceedance once every 8 to 10 seconds. The error made with this method is
dependent on the size of the length variations over time, the bigger the air bubble
plume the bigger the plume size variations. So an estimation of the error made with
this measurement is: ±10 %.
Other specific procedures about the measured parameters can be found in
paragraph 3.6.4.

The start up time is different for every measurement, but most times 1 minute is
enough to get an overall stationary situation.

3.6.3 Imposed parameters

Fixed parameters
The fixed parameters that can not be varied in the used model set-up are:
1. Approach flow tube diameter (D = 0.192 m, inner diameter)
2. Quality of the water (water from the cellar basin below the laboratory hall of
WL|Delft Hydraulics)
o Water temperature (= ±18.5ºC)
o Surface tension of the water (σ = 0.0765 N/m, is relatively high)
3. Reservoir width and depth (b * h = 1.00 m * 1.00 m)
4. ITT Flygt Submersible pump
o Type: NP 3102 MT, 100 mm
o Impeller code: 462
o Best Efficiency Point (BEP): 25 l/s with a 6.5 m water head
o Maximum water head: 12 meter

40
Variable parameters
De variable parameters are adjustable but fixed for every measurement:
1. Water depth Hw (m)
2. Fall height H (m)
3. Tube – pump distance Ltp (m)
4. Water discharge Qw (l/s)
5. Impeller revolutions (for pump air intake experiments only) (l/min)
See figure 3.1 or appendix A for a schematic overview of all parameters.

1) Water depth Hw
The water depth can vary between 0.35 and 0.95 meter. The submersible pump needs
a minimum water level of 0.24 m plus 0.11 m for a concrete pomp base. The height of
the flume walls requires a maximum reservoir water level of 0.95 meter water depth.
Measures are done only with water levels: 0.95, 0.60 and 0.40 m.

2) Fall height H
The fall height is the vertical distance between the inner side of the lower tube edge
and the reservoir water level. It is adjustable to a value of 0.98 m above the maximum
water level. That is about 1.93 m above the reservoir bottom. The minimal distance
between tube edge and the reservoir bottom is 0.40 m.

3) Tube – pump distance Ltp


The horizontal distance between tube edge and pump inlet centre line is variable with
every value between 0 to 4 meters. When the tube – pump distance is more than 3.50
meters, the pump have no effect on the air bubble plume for every possible fall height
or water discharge.
For the pump air intake measurements, the following tube – pump distances are
chosen: 0.70, 1.00 and 1.50 meters

4) Water discharge Qw
The water discharge can vary between 0 and 25 l/s. This discharge is always equal for
both pump flow and plunging jet flow. For the measurements only discharges
between 0.5 and 15.0 l/s are used. For discharges less than 0.75 l/s the jet used to
follow the tube edge curve, so results like fall distance and horizontal velocities are
affected by this phenomenon and are less useful. Besides that, the discharge gauge has
a resolution of 0.15 l/s, so exact flow measures with these small discharges are almost
impossible.
Overfall jets with discharges above 10 l/s are not stationary and stable with the
used model set-up. The reason for this is not known, but there are some possible
causes; a possibility is a non-stationary water supply from the pump, but it is also
possible that the length of the approach channel is not long enough to eliminate all big
turbulent instabilities in the flow. More about this in paragraph 4.1.

5) Pump impeller revolution rate


After the First measurements with pump air intake, it became clear that the number of
revolutions of the pump impeller is an important factor for the amount of air that is
taken in per unit time. So with equal parameters like fall height, water depth, tube –
pump distance, and even the water discharge, set to the same value with the control
valve, a higher revolution rate will give a higher pump air intake discharge.

41
In experiments without pump influence on the bubble plume, the revolution rate
of the impeller is of no importance and chosen equal to 1000 rpm to relieve pump
stresses. The maximum impeller revolution rate is 1445 rpm.
For pump air intake experiments, the revolution rate is also chosen equal to 1000 rpm.
For more information, see paragraph 4.3.

3.6.4 Measured parameters


The following parameters are measured in various experiments:
1. End depth (= an tube edge) he (m)
2. Horizontal fall distance Ljx (m)
3. Penetration depth Hp (m)
4. Bubble zone length Lb (m)
5. Maximum length and width of plunge point
6. Circular fall distance Hcir (m)
7. Air entrainment rate Qa (l/min)
8. Pump air rate Qap (l/s)

1) End depth: he (m)


The end depth is the water depth at the edge of the approach flow tube. Following the
literature of Dey (1998, 2001) there is a linear relationship between the critical water
depth (Fr = 1) and the end depth in an approach channel. The water discharge is also
calculable with only this flow parameter and the tube diameter, see paragraph 2.9.
The end depth is measured with a point gauge. Because it has a nonius or vernier
scale, it is possible to measure the water depth in 10th of millimeters. But with large
non-stable channel flows, errors up to several millimeters are inevitable.

2) Horizontal fall distance: Ljx (m)


The horizontal fall distance is the horizontal distance from the tube edge to the centre
of gravity of the plunge point.
The estimated error is ± 0.5 to 2.0 cm with discharges from 0.5 to 15 l/s.

3) Penetration depth Hp (m)


The penetration depth is the depth which the deepest air bubble in the plume crosses
once every 8 to 10 seconds.
The estimated error is 1.0 cm.

4) Bubble zone Lb (m)


The bubble zone is the horizontal distance from the tube edge to the furthest point in
the reservoir at which air bubbles, deeper than 5 cm, passes once every 8 to 10
seconds.
The minimum depth of 5 cm for air bubbles in the bubble zone is chosen because
with big air entrainment discharges of Qa ≥ 200 l/min a big horizontal flow appears at
the reservoir surface as a result of the upward circulation by the rising air bubbles in
the up going region of the plume. This surface flow can bring air bubbles up to
decimeters further from the tube edge than the length of the bubble zone, see
definition above. But this extra distance is less important in this research at which
only deep air bubbles near the pump inlet are important, so it is omitted.

42
The aeration zone is the bubble zone minus the horizontal fall distance:
La = Lb − L jx .

5) Jet cross section at plunge point


The maximum length and width of the plunge point are measured with a ruler. The
maximum length of the plunge point was not always in the middle of the jet, in case
of a U-shaped plunge, see figure 4.4. The plunge point area is of course always less
than the multiplication of the plunge point length times the width.

6) Circular fall distance Hcir (m)


The circular fall distance is the fall distance at which the jet has a more or less circular
shaped cross section. This is the result of the surface tension of the water in the jet.
Every discharge has its own circular fall distance, given the water properties and the
tube diameter. The relation between the water discharge and the circular fall distance
turn out to be linear, see paragraph 5.3.
The estimated error is ± 5.0 cm.

7) Air entrainment discharge Qa (l/min)


The air volume that entrains the water per unit of time is measured with a bubble trap
and an air flow meter. The maximum measurable air discharge is 100 l/min, with a
resolution of 1.0 l/min.

8) Pump air intake discharge Qap (l/min)


The pump air intake discharge is the total air intake volume per unit of time. This air
volume is determined with a vertical standing pipe which collects all the air bubbles
in the pressurized outflow pipe, see paragraph 3.5.
The volume of the collected air in the standing pipe is corrected for the occurring
P
pressure in the pipe by using next formula: Qap = Q% ap . The maximum estimated
Patm
error of Qap is ± 0.1 l/min.

43
Chapter 4: Qualitative observations and discussion
Observing the phenomena occurring with a plunging water jet in a sewer sump model
by the naked eye, a lot of things can be said without having measurement results. In
this chapter some observed features of plunging jets (4.1) will be described.
Following that, observations on air bubble plumes are given in paragraph (4.2). The
last paragraph (4.3) describes observed relations between the submersible pump and
bubble plumes.

4.1 Observations on plunging jets and air entrainment


After the first experiments without pump influence on the bubble plume, it turned out
that the air entrainment discharge is very sensitive to both the roughness of the jet
surface (4.1.1) and the jet cross section shape. Combined with the jet water discharge,
these are the main parameters for air entrainment.

4.1.1 Jet surface roughness


The roughness of the jet surface is one of the main important factors for the amount of
air entrainment, because a rough plunging water jet can enclose air particles in the
direct neighborhood of the plunge point. Corresponding literature can be found in
paragraph 2.3.2. Recent research from Zhu et al. (2000) demonstrates that jet surface
ripples are always needed to entrain water with plunging jets.

The jet surface roughness is at first dependent on the turbulent movements in the
approaching flow, so when this flow is more turbulent, the falling jet surface will be
less smooth. A second cause for a rough jet surface is the air friction during the fall of
the water jet. The last possible cause of jet surface ripples is the channel edge shape
and roughness.

For the model experiments with big discharges (> 7.5 l/s), the surface of the jet and
the approach flow were not stable and rough. A possible cause is a non-constant
delivery of water by the pump, but a more acceptable cause for this is an approach
channel which is not long enough to eliminate all big turbulent instabilities in the
approach flow.
In the model set-up, the approach flow channel must be able to lower below the
flume walls. So it was needed to have a vertical connection at the upstream end of the
approach channel, see figure 4.1. But the result of this connection was a highly
turbulent and rough flow at the upstream end of the approach flow. It is plausible that
the channel was not long enough to get a steady and smooth outflow.

44
Figure 4.1: Connection upstream end of the approach flow channel

In literature it is described when the nozzle length is 50 times the diameter; the
flow in the nozzle is completely developed and will not change with longer nozzles.
So the rule ‘50 times the diameter’ is probably also needed for tube channels. In that
case the tube must be LT = 50 * 0.20 = 10 meter. But a 10 meter long tube is hard to
handle, especially when the height of the tube must be changed often. So a length of
25 times the diameter (= 5.0 m) was chosen to get a more manageable model set-up.
When the length of the tube is divided by the hydraulic diameter at critical
 Ac 
flow  DH ,c = 4  instead of the tube diameter D, what is reasonable with
 0.5δ c D 
channel flows, the dimensionless length of the tube is than Lˆ = L D ≥ 50 for water
T T H

discharges Qw ≤ 2.4 l s , see also figure 4.2. So it can be assumed that discharges less
than 2.4 l/s are completely developed, but discharges to 5 l/s were also visibly stable
and stationary. In future measurements it is recommended that approach channels for
overfalling jets are at least 35 times the hydraulic diameter to get a stationary
situation.
To improve the situation for bigger discharges than 5.0 l/s, some measures are
taken. First a trouser leg is implemented at the upstream part of the approach tube. It
is like a sac around the water flow which enlarges the local roughness, but the effect
was small. So a sharp weir is implemented at the upstream end of the approach tube.
This measure allowed the water flow to be sub-critical, but still the instabilities in the
flow are not vanished. Instable flows with big discharges are accepted then.

45
Figure 4.2: Dimensionless length of the tube at critical flow

Approach channel edge


The approach channel edge is an important factor for the roughness of the falling jet.
Because the ripples on the jet surface are the cause of the amount of air entrainment,
the channel edge is one of the causes of ripples on the falling jet surface. So the
smoother and sharper the channel edge, the less ripples are formed on the jet surface.
This means that air entrainment in real sewer sumps are difficult to model with a
laboratory set-up.

In the used model set-up the approach tube edge is sharp and has no big
irregularities. But especially at small discharges, a part of the jet flow is branched off
at the channel edge. It is like the flow sticks to the channel edge. Possibly the surface
tension between the water and the PVC tube is the cause of this effect, but maybe also
the Coãnda-effect plays a role in this situation. The Coãnda-effect is the bending of a
flow along a convex surface, see figure 4.3 for an example.

Figure 4.3: Schematic demonstration of the Coandă effect (Pol-O.R. Bear, 2006)

46
4.1.2 Shape of the plunge point
The shape of the jet cross section is an important parameter with air entrainment,
because air will only entrain when a small air pocket in the direct neighborhood of the
plunge point is enclosed by water. The air bubble is then transported into the reservoir
by the submerged jet. No air bubbles are entering into the falling jet itself, the jet will
at most disintegrate in separate water packages and droplets, dependent on the fall
height, the turbulent movements in the falling jet and the viscosity of the water.
When the perimeter length of the plunge point increases, more air has a chance of
getting enclosed by water and more air will entrain the reservoir. So a stretched jet
entrains more air than a massive circular jet of the same discharge, fall height, etc.

It is clear that a falling jet from a horizontal open approach channel has a specific
characteristic shape. This shape is dependent of the rate of filling of the tube and has
for any fall height a different horizontal cross section in the plunge point. The plunge
point shape is in common U-shaped with relatively small fall heights; this becomes
more and more circular with larger fall heights.

4.1.3 Features which influence the jet shape during the fall
When a water jet leaves the approach channel the shape of the jet cross section
changes during the fall. There are four features that influence the jet cross section
shape:
• Acceleration of the water jet by gravitational forces
• Velocity profile of the channel flow at the channel edge
• Turbulent movements in the jet
• Surface tension forces in the water jet

Gravitational acceleration
Because falling water accelerates by gravitational forces, the cross sectional area of
the jet in the plunge point is smaller than the cross sectional area at the channel end.
This is valid only when the net water jet cross section is taken into account, so air
spaces between water particles are omitted.
Eventually when the fall height is high enough, a falling jet will be disrupted
when turbulent movements in the jet are strong enough. In case of a coherent stable
jet, the jet is stretched until water surface tensions become dominant and the jet is torn
up into droplets. The breaking up of the falling jet has a large influence on the air
entrainment process at the plunge point. More air will entrain with a disrupted
plunging jet and also the penetration depth will decrease, see more about this in
paragraph 4.2.3.

Velocity profile of the channel flow


The horizontal velocity profile in a circular channel is not uniform over the flow
width. Because of that, some flow parts fall further from the tube edge than other
parts and so the jet cross section shape will change during the fall.

47
Figure 4.4: Turbulent velocity profile in circular open channel flow

In an open tube flow, the parts A and C of the flow cross section, see figure 4.4, has a
lower mean velocity than part B. That is the effect of the tube wall friction. So part A
and C will reach a smaller horizontal distance than the middle part of the jet. The
horizontal cross section of the jet becomes U-shaped with the convex part to the front.
See figure 4.5 for a development of the jet cross section shape during the fall.

Figure 4.5: Development of the horizontal jet cross section during the fall.

This U-shaped jet cross section occurs always with fall heights H = Hu < Hcir, with
Hcir and Hu the fall heights by which the jet has a circular respectively U-shaped jet
cross section. A U-shaped plunge point turns out to be very important for air
entrainment, see paragraph 4.1.4.

Turbulent movements in the falling jet


There are two causes for turbulent movements in falling jets. The first one is the
turbulence that is already in the flow and that is the result of the velocity gradients in
the approach channel flow, due to the tube wall frictions. The second cause is the
turbulence originated from the air friction during the fall of the water jet. These
turbulent movements are the cause of ripples on the jet surface and, in case of heavy
turbulence, the deformation or disruption of the falling jet. So turbulence in the falling
jet is also an important factor for air entrainment with plunging jets

48
Surface tension of the jet water
The surface tension of the water in the jet will try to minimize the length of the jet
cross section perimeter. So the surface tension force tries to make the shape of the jet
cross section as circular as possible, see also figures 4.5 and 4.6.
No further research is done on the effect of different surface tensions of plunging
water jets.

Figure 4.6: Deformation of jet cross section; the jet is seen from behind

4.1.4 Plunge point enclosing funnels


When a jet plunges into a reservoir, the reservoir water directly around the plunge
point will be pulled into the reservoir with the submerged jet. At the location where
these two water surfaces meet, emerges a funnel or trumpet around the plunge point,
see figure 4.7. In this funnel air can be pulled into the reservoir due to irregularities at
one or both water surfaces, see paragraph 4.1.1. So in other words, the better the
reservoir surface water is able to adjoin the shooting water jet, the less air will be
entrained with plunging jets.

Figure 4.7: Sketch of a funnel around a water jet in the plunging point

49
In some cases the approaching water at the reservoir surface is hampered a bit to
flow to the plunge point, for example when the plunge point is in the neighborhood of
a wall or floating obstacle. This has directly influence on the local funnel depth
because of the reduced approaching reservoir surface water. The funnel depth
increases and thus air can be entrapped easier within this deep funnel, so the amount
of entraining air increases also.

In case of a U-shaped jet, reservoir surface water cannot adjoin the shooting jet in
the concave part of the U-shaped plunge point as easy as at the convex part. So, the
funnel depth at the concave part is deeper and entrains far more air than the convex
part of the plunge point.

Also with air entrainment measurements, the bubble trap wall must not be placed
very close to the plunge point to prevent funnel depth increase and air entrainment
manipulation. Bubble traps which enclose plunging jets with a gap, like in figure 4.8,
are not reliable for entrained air volume measurements. It is better to have a bubble
trap without a gap, also then practically all entrained air bubbles will be captured.

Figure 4.8: Not reliable bubble trap design

4.1.5 Jet plunge angle


The effect of different jet plunge angles cannot be simply measured with channel
overfall jets. The angle is the result of the combination of the water discharge and the
fall height. So the plunge angle is not independent and can therefore not be altered
without also changing other parameters. The following statements are from literature
and are done for fast and small water jets from nozzles.

The amount of air entrainment increases with smaller jet plunge angles. The most
probable reason for this is the small angle between both jet and reservoir water
surfaces, so air can be enclosed easier than with vertical jets. This is of course only
valid for the lower side of the plunging jet. As a matter of fact, Sande & Smith (1973)
say it has only significant effect with plunge angles αj < 60°, which is not the case for
the experiments done with the used model set-up.
No further investigations are made on this aspect.

50
4.1.6 Approach flow channel inclination
When the approach flow channel makes a slight down hill (positive) inclination of
1:400, there are already some visible changes of the falling jet. The horizontal
velocity of the flow is higher and the falling jet surface is rougher.
The extra positive acceleration of the flow, at a slight positive channel inclination, is
almost not slowed down by the small wall surface roughness of the PVC approach
tube. So that is probably the reason why the horizontal velocities of the jet are then
visibly higher. The roughness of jet surface is than the result of the higher flow
velocities and velocity gradients in the approaching channel flow.
An unintentional up hill (negative) inclination of the approach flow channel has
the opposite effect. The horizontal velocity of the flow is reduced and the falling
water jet surface was visibly smoother.
Literature confirms the above-mentioned findings for the horizontal end velocity
of the approach flow, discussed in paragraph 2.9.3.

These findings are only observed qualitatively and no extra measurements are
done to explore these results. But it is an advice that the approach flow tube must be
leveled accurately. This may be the cause for some minor unexpected deviation in the
results.

4.2 Observations on air bubbles and air bubble plumes


4.2.1 Transport of air into the reservoir
When air bubbles are formed at the plunge point of falling jets, the submerged jet will
transport the air bubbles deeper into the reservoir. Dependent on the air bubble sizes
and the angle of the submerged jet, are the bubbles able to leave the submerged jet
sooner or later. It is also possible that the submerged jet (from now on called just
‘jet’) disintegrates by hitting an obstacle and the bubbles can leave the jet. When the
bubbles have left the jet and no other flow is transporting them further, the bubbles
will rise up to the water surface and leave the reservoir. So the air bubble plume is
created. See figure 4.9 for an impression.

Figure 4.9: Impression of a bubble plume (the reservoir 1.0 m deep)

51
4.2.2 Air bubble properties
Bubble size and water surface tension
The size of the air bubbles in an air bubble plume is relatively constant, because this
is mostly dependent on the water surface tension and the intensity of the shear stresses
in the reservoir. The bigger the water surface tension, the stronger the air bubbles are
and thus the better the air bubbles are able to resist the occurring shear stresses. When
an air bubble is relatively large to the prevailing shear stresses, the air bubble breaks
up into smaller ones. But the same turbulent movements also cause collisions between
air bubbles, which can merge together and move on as one. These two effects are the
causes that air bubbles in a bubble plume do have more or less the same size. This is
also visually observed.
The relation between those two effects are combined in the Weber number
V2
We = , which is proportional to the ratio of the inertial force to capillary force
σ ρL
(i.e. surface tension).

The used water in the model has a relatively high surface tension. With a capillary
tube with a radius of r = 0.2 mm, is measured a capillary rise of hm = 78 mm at ±18ºC.
This is pretty high in comparison to the measured capillary rise of tap water: ht = 76
mm also at ±18ºC. So the accompanying water surface tensions are respectively: σm =
0.0765 N/m and σt = 0.0746 N/m, with both ρw = 1000 kg/m3. The high surface
tension in the model reservoir is probably the result of the salty minerals that are
present in the cellar basin of the laboratory hall of WL|Delft Hydraulics. Therefore it
can be assumed that the air bubbles in the used model water are a bit larger than with
clean tap water.

The water in real sewer sumps is strongly polluted with surfactants, so the surface
tension in real sewer sumps will be much lower than in the model set-up. Air bubble
sizes are then (much) smaller and that affects also the air bubbles rise velocities. The
residence time in the reservoir water will be longer which enlarges the chance for air
bubbles to be taken up by the submersible pump, with all negative side effects. So
results of the used model set-up are thus too promisingly for real sewer sumps and
cannot be used as design rules unconditionally. First the effects of other water
properties must be examined carefully.

Air bubble residence and penetration depth


The mean reservoir air bubble length of stay is mainly dependent on the rise velocity
and the penetration depth. When an air bubble is relatively large, it can easier break
away from the jet and will have a smaller penetration depth. The big bubbles have
also a big rise velocity which brings them quickly to the water surface. So the sizes of
the bubbles affects the residence time in two ways.
As described above the water surface tension is mostly the cause of the bubble sizes.
So with a bigger water surface tension the penetration depth will be bigger and the
length of stay shorter.

The angle of the submerged jet is also important for the air bubbles residence
time, because with a oblique jet, the rising force on the air bubbles is not opposite to
the jet momentum but under an angle. So air bubbles can relatively easy escape the jet
upwards. So the penetration depth is much less with oblique jets (αj ≤ 70º).

52
The relative low density of the air bubble plume is positive for the rise velocity of
the individual air bubbles, that is because the plume rises as a whole and causes a
reservoir circulation. So not only the air bubbles rise, but also the water in between.
This means that the denser the air bubble plume is, the faster the air bubbles rise. The
rise velocity of the air bubbles in the used model is about 0.15 to 0.20 m/s. These
values are not measured exactly.

Remedy for a deep penetration


A simple way to shorten the residence time of the air bubbles and to decrease the
chance of an air bubble to be taken in at the pump is to dissipate the submerged jet by
hitting an object like a horizontal rigid plate, see figure 4.10. When the plate below
the water level forces the jet to flow horizontally, air bubbles cannot be brought
deeper into the reservoir. Besides that air bubbles are not hindered any more to rise up
to the water surface. Recommending dimensions for the rigid plate are 2 times the
approach flow channel diameter D wide, 1 D below the lowest reservoir water level
and 2.5 to 4 D long, dependent on the location of the plunge point and the angle of the
submerged jet. These lengths are defined by gained experiences with the model set-up
and analysis of the measured data.

Figure 4.10: Schematic drawing of an air bubble plume with or without an obstacle

4.2.3 Submerged jet obstruction


Sometimes the transport of entrained air bubbles underneath the plunge point of a
near vertical plunging jet (αj = 75-90º) is not big enough. The air bubbles are not
transported deeper into the reservoir and form a blockage for the submerged jet. The
jet loses its momentum by dissipation against the congestion of self entrained air
bubbles. So air bubbles are brought less deep into the reservoir and the blockage of
the submerged jet becomes bigger until equilibrium is reached. See figures 4.11.a and
4.11.b for the penetration depths of an obstructed and a non-obstructed submerged jet.

In case of a more oblique jet (αj ≤ 75º) the resulting horizontal force of the jet on
the bubble plume makes sure that the entrained air bubbles will flow away sideways.
So the more the plunging jet is vertical the more the entrained air bubbles are able to
block the submerged jet and the less the penetration depth will be.
This phenomenon is the result of the rate between the jet momentum and the
rising force of the entrained air bubble plume along with the jet plunge angle. So it
can be said that a big, fast, circular jet with a smooth surface has almost no chance of
getting obstructed by entrained air bubbles and so the penetration depth will be very

53
big. Most times this is the case for plunging jets with relative small fall heights (0.20 -
0.40 m) and stable but big enough discharges (1.5 - 3.0 l/s. With smaller fall heights
the jet plunge angle is not vertical enough to get a deep penetration. Jets from bigger
fall heights have more momentum but have also rougher surfaces, so more obstructing
air bubbles will be entrained. This is also the case for bigger discharges then 5.0 l/s,
see also paragraph 4.1.1. Smaller discharges do not have a big enough momentum for
a deep penetration.
The biggest jet obstructions with entrained air bubbles are from jets with high fall
heights (≥ 0.60 m) and small discharges (≤ 1.0 l/s). These jets are relatively rough and
therefore entrain a lot of air. With these small discharges the jet has almost no
momentum for breaking up blockages and plunges also near vertically, which is
needed for jet obstruction.

(a) (b)
Figure 4.11: Pictures of two penetration depths with a slight different discharge and fall height

4.2.4 Reservoir circulation


In case of a big air entrainment discharge, the bubble plume causes a circulation in the
reservoir. At the location of the rising bubbles the water is flowing upwards and at the
reservoir walls and the plunge point the water is sinking. The bigger the bubble plume
with respect to the reservoir dimensions, the bigger the downward flow at the walls.
At discharges Qw ≥ 7.5 l/s, the downward flow at the reservoir walls were that strong
that they could trap air bubbles into the reservoir. These air bubbles were just floating
more or less horizontally against the reservoir walls, see also figure 4.12.

The up going flow in the middle of the reservoir is flowing against the downward
flowing submerged jet. This feature decreases the penetration depth of the bubble
plume, but the reservoir depth was just not deep enough at these discharges to
measure this effect.

54
Figure 4.12: Floating air bubbles against the reservoir walls due to reservoir circulation

4.3 Observations on pump air intake


From experiments with pump air intake experiments the following observations are
made. Air bubbles are only taken in by the pump when they are in the direct
neighborhood of the pump inlet. The influence sphere of the pump is for discharges
Qw ≤ 15.0 not more than a radius of 0.10 m around the pump inlet centre.

The most striking finding on the intake of air at the submerged pump was that the
revolution rate of the pump impeller was very important. At some impeller revolution
rates a circulation in front of the pump inlet emerges, this is called a pre-rotation. It is
known that a pre-rotation appears when a pump is not pumping in its Best Efficiency
Point (B.E.P.), so the revolution rate times the impeller pitch is not equal to the
occurring water discharge. When the revolution rate is too high, the water in front of
the inlet spins in the same direction as the impeller, in case of a too low revolution
rate the pre-rotation is in the counter direction. Figure 4.13 shows a pre-rotation
vortex.

Figure 4.13: Pre-rotation vortex at the pump inlet

It is likely plausible that a pump in its B.E.P. has no influence on the water in
front of the pump inlet because of the lack of pre-rotation. So it was better to have no
pre-rotation at all, but unfortunately it was impossible to pump small water discharges
(0.5 - 5.0 l/s) with the accompanying low revolution rates to keep the pump in its
B.E.P. A compromise is an impeller revolution rate of 1000 rpm, this is the lowest
revolution rate at what the pump can deliver the requested water discharges (0.5 –
15.0 l/s) with the highest fall height (Hmax = 0.98 m).
More on this provided in section 5.7.

55
The pre-rotation of the pump was able to generate an extra reservoir circulation.
When the reservoir dimensions where small, about ≤ 2.0 m3, a vortex was formed
around the plunge point. Because of this vortex the funnel around the plunge point
becomes much deeper and that certainly influences the air entrainment. Unfortunately
the bubble trap was far too large to measure the total air entrainment in such a small
reservoir and so it was not possible to do those extra measurements.

Sometimes when the pump air discharge Qap is high compared to the water
discharge Qw the pump can not pump away the air inside the pump casing. The pump
impeller is then spinning in an air pocket and the pump pumps no more water. This
Qap
phenomenon occurs with ratios ≥ 0.01 , especially with low water discharges.
Qw

The pump impeller revolution rate is of no importance with air bubble plume
measurements without the influence of the pump. The revolution rate is then also
chosen at 1000 rpm to relieve pump stresses.

56
Chapter 5: Results and analysis
In this chapter is described the measured data and conclusions that are visible from
them. In paragraph (5.1) the end depth results are shown and are compared with
empirical formulas suggested in literature. In paragraph (5.2) is depicted the modeling
of the horizontal fall distance of a free overfalling jet. In paragraph (5.3) is described
the modeling of the circular fall height and in paragraph (5.4) is shown an empirical
relation for the aeration length. A comparison with a formula in literature appeared to
be impossible. The penetration depth and the relation with the air entrainment are in
paragraph (5.5). The relation between discharge and the air entrainment discharge are
given in paragraph (5.6). In the last two paragraphs is told something about the pump
impeller revolution rate (5.7) and the pump air discharge (5.8).

5.1 End depth measurements


The end depth he, see Appendix A, is measured a couple of times in experiments
where only parameters changed which had no effect on the end depth. Figure 5.1
shows the results.

Figure 5.1: Measurement series of dimensionless end depth

The spreading of the end depths is very small for dimensionless discharges
ˆ
Qw ≤ 0.15 (Qw ≤ 7.5 l/s). For larger discharges there is a significant spread in the
results, which is probably due to the unstable approach channel flow, see paragraph
4.1.1.

Measurement 5 was done with a trouser leg as stilling measure at the upstream
end of the approach channel. This has some effect on the stability of the approach
flow so this measurement is assumed to be the best end depth measurement. To

57
11
Qw  h 6
compare the results of measurement 5 with the formula = C  e  written in
g D 2.5 D
literature in paragraph 2.9.4, figure 5.2 is given.

Figure 5.2: Comparison of various empirical formulas and the measured data with error bands

The error bands are given for the measured data (∆x = ±0.003 ≈ 0.15 l/s). The
error for in the y-axis is negligible in the figure when the estimated error is 0.001 m.
The results of the measured data come very close to the formula with constant C =
1.54, which is the formula of Rajaratnam & Muralidhar (1964).

5.2 Horizontal fall distance


When the water discharge, the end depth and the horizontal fall distance are known, it
is possible to calculate the fall curve of the plunging jet. The horizontal flow velocity
Q
at the tube end is equal to: ve = w . The air friction forces are supposed to be zero, so
Ae
the fall time is known by dividing the fall distance with the horizontal flow velocity:
L jx
Tj = . Because these fall times are not exactly equal for different discharges with
ve
2
the same fall height, an extra height is added to the fall height: H e = H + he , with he
3
the end depth. This extra fall height is the location of the mass centre of the flow at
the channel edge.

An extra vertical velocity at the channel end we is added, by knowing the flow is
at that location strongly curved. This vertical velocity is calculated by using equation:
1
H e + weT j − gT j2 = 0 . Every fall height He has a different fall distance Lj,x, but the
2
vertical end velocity we remains the same, so the average vertical velocities we are
calculated to determine the theoretical curves. The curves are represented by the

58
 1 
formula: ( x(t ); y (t ) ) =  vet ; H e + wet − gt 2  and they are a good approximation of
 2 
the measured data see figure 5.3.

Figure 5.3: Horizontal fall distances with various discharges and the accompanying theoretical
trajectories of the jet axes

Figure 5.4 gives an overview of the various average flow velocities at the
beginning of the jet for various discharges. The average vertical flow velocity we at a
discharge of Qw = 0.5 l/s differs much from the we values of other discharges. That is
probably due to the occurrence of the surface tension or the so called Coandă effect,
see also paragraph 4.1.1. Because of this, a part of the horizontal velocity will be
transferred to the vertical velocity, so the horizontal velocity is calculated too big and
thus is the fall time assumed too small. The initial vertical velocity must now be large
enough to reach the vertical fall distance in that short fall time.
In figure 5.3 can be seen the result of this effect on the theoretical curve for a
discharge of Qw = 0.5 l/s. It does not fit the measured data as good as the rest of the
theoretical curves. In fact the inclination of the curve at Ljx = 0 m should be more
horizontal and the rest of the parabolic curve should be more curved.

59
Figure 5.4: The calculated average flow velocities at the channel end for various discharges

5.3 Circular jet cross section fall height


As described in paragraph 4.1.3 is the shape of the plunging jet cross section circular
and massive, when falling a certain distance. This circular fall height is measured for
a couple of discharges, as can be seen in figure 5.5.

Figure 5.5: Dimensionless fall height when the jet cross section is most circular and massive

The chart is given in a dimensionless form, because probably this relationship is


also valid for other approach flow diameters. In case of a smaller approach channel
diameter, the fall length to deform an overfalling jet with the same dimensionless

60
discharge is also smaller. But there are no measurements done with other approach
channel diameters to check the dimensionless relationship.

In spite of the relative big error margins for the fall height and the discharge, a
linear relationship in figure 5.5 between the fall height and the discharge seems in this
range fairly possible. The dimensionless formula of the fitted line is: Hˆ cir = 102.0Qˆ w
( H cir = 0.39Qw ) . From now on red circles in other charts mark this relationship, to see
if the circular and massive jet cross section has any influence on the shown results.
See for example figure 5.6.

5.4 Aeration length


The aeration length La is the length of the bubble zone Lb minus the horizontal fall
distance Ljx. The relation between the aeration length and fall height is shown in
figure 5.6. There is almost no relationship visible between those two parameters,
which is on first sight not plausible. Only an effect of the circular fall height is
somewhat visible, see the red circles in the chart. Jets with a U-shaped plunge point
(lower side from the red circles on the curves) have a bit shorter aeration lengths than
jets with a circular shaped plunge point (upper side of the red circles).

Figure 5.6: Aeration length from plunging jets with various water discharges and fall heights

The sometimes parallel deviations in the curves may be the effect of a slight
approach channel inclination at some fall heights. Especially at the fall height values
H = 0.20, 0.30 and 0.40 m, the inclination was measured less accurately.

Figure 5.6 shows that the aeration length is mainly dependent on the water
discharge. When all results of the aeration length are plotted against the water
discharge, the chart in figure 5.7 is obtained.

61
Figure 5.7: Determination of the maximum aeration length versus discharge

The best fitting line though all aeration length results is: La = 0.37Qw0.58 and the
best fitted formula for the maximum aeration length is: La ,max = 0.5Qw0.5 . The constants
s
0.37 and 0.5 are not dimensionless, they have the dimension of . Figure 5.8
m
shows that all measured aeration lengths are between the following relations:
0.3Qw0.5 ≤ La ≤ 0.5Qw0.5 .

La
Figure 5.8: Relation between Qw
and fall height H

62
Comparison of measured data with formula from literature
Because the bubble zone length formula described by Nakasone (1987)
0.134
Lb = 0.0629 ( H j + 1.5hc ) qw0.666 ,
see also paragraph 2.8.3, is only valid for free overfall nappes in stead of free overfall
jets from circular channels, a comparison with the measured data in this experiment is
not without problems. The difficulty is that an overfall nappe is 2D and an overfall jet
is a 3D phenomenon, so not only the plunge but also the flow in the reservoir is
different because of the third dimension.
By just filling in the measured results in the formula of Nakasone (1987), figure 5.9 is
Q
obtained. With this formula is approximated qw = w and is La = Lb – Ljx.
D

Figure 5.9: Comparison of nappe jet aeration length formula of Nakasone (1987) with measured
data

It is visible that the nappe jet formula is not applicable for jets from circular open
channels. The formula of Nakasone gives results that are about five times bigger then
the measured data of aeration lengths. The formula of Nakasone has a small fall
height dependency, which is not found in the measured results.

Aeration length in different water depths


There are some experiments done with different water depths (but with equal fall
heights). The results are shown in figure 5.10.

63
Figure 5.10: Aeration lengths for various water depths, fall heights and discharges

The figure shows a little effect of the water depth on the aeration length. The
aeration length decreases (0 – 40 %) with less deep water depths, especially for water
discharges between 3.0 and 10.0 l/s. But there is no clear relationship visible.

5.5 Penetration depth


The next surprising results are the measured data for the penetration depth, see figures
5.11 and 5.12.

Figure 5.11: Measured data for the penetration depth for various discharges

64
Figure 5.12: Measured data for the penetration depth for small discharges

At first the penetration depth is clearly not constant for any discharge and/or fall
height. So the rule of thumb given in literature that the penetration depth is more or
less equal to 2/3rd of the fall height is not correct at all. In fact, jets with the smallest
fall heights have mostly the biggest penetration depths.
For the minimum discharge measured, 0.5 l/s, the penetration depth is almost
constant for any fall height. But when the discharge increases to 2.0 l/s, the
penetration depth increases surprisingly faster for low fall heights than for jets from
high fall heights, see also figure 5.12 for a zoomed in version of figure 5.11.
The horizontal lines at Hp = ± 0.94 m represents the bottom of the reservoir, so the
reservoir was unfortunately not deep enough to measure the whole range of interest.

At high discharges (Qw > 2.0 l/s) the penetration depth decreases at some
apparently arbitrary fall heights. There is no unambiguous explanation for this
phenomenon. Maybe this is the result of a slight approach flow inclination at some
measurements, but that is uncertain.

In figure 5.12 the red circles indicates the discharge-fall height combination where
the jet is massive and circular at the plunge point. So the plunging jet entrains then
relative small amounts of air into the reservoir. The relation in paragraph 5.3 is used
for determining the combinations between circular discharge and fall height.
On the left side of the red circles on the curves, the jet plunge point is U-shaped
and the right side of the curve, the jets cross section is more or less circular shaped. It
is clear that the penetration depth increases much when the jets cross section is near
massive and circle shaped. This is visible too in figure 5.13, where the discharge Qw is
divided by the circular discharge Qcir for that specific fall height.

65
Qw
Figure 5.13: Relationship between penetration depth and
Q cir

In contrast to the charts in figures 5.11 and 5.12 it is clear that bigger fall heights
penetrate deeper into the reservoir than small fall heights when we look at the ratio
Qw
in stead of Qw. Jets from big fall heights do not need to be as circular in the
Qcir
plunge point as jets from small fall heights to penetrate to the same depth.
At last figure 5.13 makes clear that the penetration depth increases rapidly when
the discharge Qw increases already from ¾th of the circular discharge Qcir.

Following the literature in paragraph 2.8.2, the penetration depth is proportional to


the nozzle or jet diameter with jets from nozzles. But with these experiments no
unambiguous nozzle or jet diameter is available, so for this case the end depth is
assumed to be the proportional parameter for the penetration depth. The penetration
depth Hp can be made dimensionless by dividing with the end depth he and figure 5.14
appears.
Q
This figure is only useful for values of about w < 0.85 , because from that point
Qcir
on some air bubble plumes begin to reach the bottom which affect the dimensionless
penetration values, see figure 5.13. From the figure can be concluded that the
Hp
dimensionless penetration depth is more or less constant ≈ 27 for
he
Q Q
0.20 < w < 0.60 and for 0.60 < w < 0.85 the dimensionless penetration depth
Qcir Qcir
Hp
increases from 27 to about 33.
he

66
Figure 5.14: Dimensionless penetration depth with various fall heights

Qa
The effect of the entrainment ratio on the penetration depth is examined. The
Qw
results are given in figure 5.15.

Qa
Figure 5.15: Relation between penetration depth and entrainment rate
Qw

Figure 5.15 makes clear that a negative relation between the penetration depth and
the entrainment rate exists, except for the horizontal curve parts which represents the
reservoir bottom, all curves have a positive slope and so, the lower the entrainment

67
ratio, the deeper the penetration depth. This relation is also described in paragraph
4.2.3.
In the figure it is indicated by red circles, at which discharges the jet plunge points
are circular. It is recognizable that the penetration depth is maximal and the
entrainment rate is minimal for circular discharges, especially at the fall height curve
of H = 0.60 m

The penetration depth formula of Sande & Smith (1975) which includes the
buoyancy forces of the entrained air bubbles see paragraph 2.8.2, is interesting to look
at. The formula is valid for only vertical jets from nozzles, but still a comparison is
done with free overfall jets. See figure 5.16 for the calculated penetration depths.

Figure 5.16: Calculated penetration depths from the formula of Sande & Smith (1975) for
vertical plunging jets with measured data

After comparing the calculated penetration depth with the measured penetration
depth in figures 5.11 and 5.12, the following things can be remarked: The penetration
depth is bigger for small fall heights than for big fall heights. This is the result of the
air entrainment discharge in the formula and corresponds to the measured data. In
both results, the penetration depth decreases with higher discharges, but the decrease
of the penetration depth is higher for the calculated penetration depths. That is
probably the result of the inclination of the measured plunging jets, which is not taken
into account in the formula, this way air escapes easier sideways underneath the
plunge point and the air bubbles can be brought a little deeper into the reservoir than
at vertical jets.
Figure 5.16 shows that a penetration depth of 3.0 m is possible with these
parameters. With a maximum possible penetration depth of 0.95 m in the model set-
up, this can not be confirmed nor denied. Scientifically it is interesting to perform
more research on this subject.

68
5.6 Air entrainment discharge
De results of the air entrainment discharge measurements with the bubble trap are
given in figures 5.17 and 5.18.

Figure 5.17: Air entrainment with various fall heights

It can be concluded that as well as the fall height as the discharge has a positive
relation with the air entrainment, but a decreasing plunge point perimeter can
counteract these effects. So the shape of the cross section profile of the plunging jet
has a strong influence on the amount of air entrainment. On the right side of the red
circles on the curves the plunge point is U-shaped and on the left side the plunge point
is circular shaped. Figure 5.18 makes even clearer that the water discharge Qw with a
Q
minimal entrainment ratio a is equal to the circular discharge (Qw = Qcir).
Qw

Unfortunately it is impossible to measure more than 100 l/min entrained air with
the used bubble trap and air flow meter. It is fairly possible to entrain more air with
the used model set-up. Besides that, air entrainment discharges in the order of 10-2 –
100 l/min are possible too with this model set-up, but the used air discharge gauge was
not able to measure them.

69
Figure 5.18: Relation between entrainment ratio
Qa and ratio Q w
Qw Qcir

Relation between jet plunge velocity and air entrainment discharge


Unless the plunge velocity range is considerable in the experiments done, the effect of
the total jet velocity in the plunge point is clearly not the main cause of the amount of
entrained air, see figure 5.19. That is in contrast to the statement given in literature,
see paragraph 2.6.2, although in accordance with the findings of Zhu et al. (2000), see
paragraph 2.3.4. The minimum entrainment increases with higher plunge velocities,
but that can also be the effect of the bigger jet surface irregularities which are the
result of the higher fall heights.

Figure 5.19: Relationship between air entrainment discharge and the total plunge velocity

70
Comparison measured data with literature formula for Qa
The by Sande & Smith (1976) mentioned aeration length formula Qa = 0.015 X 0.75 ,
−1.5
see paragraph 2.6.2, is valid for values for X = d N2 V j3 L0.5
j (sin α j )

between X = 10−4 − 10−2 . Their experiments were carried out with jets produced from
nozzles with dN = 2.85 – 10 mm, lN/dN = 50, Vj = 2 – 5 m/s, Lj < 0.5 m/s and αj = 20 -
60º.
The experiments done in for this thesis have complete other values for X:
X = 10−1 − 102 . Some other parameters in these experiments differ from the
experiments of Sande & Smith too: The imaginary channel end diameter is
4 Ae
approximated as the nozzle diameter: d%e = d N and is between 30 – 120 mm,
π
Lj = 0.2 – 1.2 m and αj = 60 - 82º. When using these parameters in the formula of
Sande & Smith, figure 5.20 appears.
It is clear that the calculated values as well as the shape of the curves are very
different to the experimental results given in figure 5.17. Therefore the formula of
Sande & Smith is useless for predicting entrainment discharges with plunging jets
from circular open channels.

Figure 5.20: Chart to compare air entrainment discharge calculated with formula of Sande &
Smith (1976) and measured data in figure 5.17

Qa
Comparison measured data with literature formula for
Qw
Following Chanson (1994) the dimensionless air entrainment is proportional to
Qa
= C * Fr 2 for Vj < 5 m/s, see paragraph 2.7.1.
Qw
With the following estimations V0 = 0 m/s and d j ≅ d e , figure 5.21 results.

71
Figure 5.21: Relation between entrainment ratio and squared Froude number

The first thing that is noticeable from figure 5.21, there is no such simple
relationship between the entrainment ratio from overfalling jets and the squared
Froude number, mentioned by Chanson (1994) for jets from nozzles. But there are a
couple of things that are worthy to remark: Again the entrainment ratio is minimal
when the plunge point of the jet is close to circular. In case of Qw < Qcir (right side
from the red circles on the curves) and the squared Froude number Fr2 < 26, the
entrainment ratio is about zero. When the squared Froude number Fr2 > 26 (and Qw <
Qcir) a relationship between the entrainment ratio and the squared Froude number can
Q
be approximated by: a = 0.025 ( Fr 2 − 26 ) .
Qw
More research is needed to confirm this relationship.
When Qw > Qcir (left side from the red circles on the curves) the entrainment ratio
increases with decreasing squared Froude number. In these parts of the curves the
plunge point is U-shaped and entrains relatively big amounts of air into the reservoir.
No relation is formulated to predict these entrainment ratios.

5.7 Pump impeller revolution rate


The most striking finding on the intake of air at the submerged pump is the
importance of the pump impeller revolution rate. When all parameters are set the
same, even the discharge, the pump air intake discharge vary sometimes more than 10
times, see figure 5.22.

72
Figure 5.22: Effect of the impeller revolution rate on the air discharge taken in by the pump

Air is pushed away from the pump inlet when the pump is not pumping in its Best
Efficiency Point (B.E.P.) that is because a vortex is formed just in front of the pump
inlet. The pump impeller is not designed for pumping low discharges 0.5 to 5.0 l/s
with high water heads, so the revolution rate is mostly far away from its B.E.P., see
figure 5.22. If the impeller revolution rate was not more than twice the B.E.P.
revolution rate, the impact on the amount of pump air intake (Qap) was not very high,
but if the revolution rate becomes more than twice the B.E.P., the pre-rotation was
strong enough to push away almost all air bubbles near the pump inlet.

5.8 Pump air discharge


The following three figures (5.23, 5.24 and 5.25) show the pump air discharges for
various water depths Hw, tube – pump distances Ltp and fall heights H.

Figure 5.23: Pump air discharge with water depth Hw = 0.40 m and rpm = 1000

73
Figure 5.24: Pump air discharge with water depth Hw = 0.60m and rpm = 1000

Figure 5.25: Pump air discharge with water depth Hw = 0.84m and rpm = 1000

The first thing that can be said about these figures is that the pump air discharge
does not differ much between water depths H = 0.40 m and H = 0.60 m. That is
because the penetration depth is mostly more than 0.60 m and so the bubble density at
the reservoir bottom (the location of the pump inlet) is somewhat the same. Another
reason is the more or less constant aeration length with varying fall heights and water
depths, see respectively figures 5.6 and 5.10, so the shape of the bubble plume is
similar too.

74
The curves in figures 5.23 and 5.24 are very steep. The vertical axis is logarithmic
which makes the curve in the upper part even steeper. The conclusion from this is: air
bubbles stay pretty close to each other and the transition between water with or
without bubbles is relatively sudden, in particular with small bubble plumes. The air
bubbles in the plume have almost constant sizes so the bubbles have almost the same
rise velocity and path.

The curves in figure 5.25 are a bit different from the foregoing two. That is the
result of the increased importance of the water depth. With this water depth (Hw =
0.84 m) the air bubbles at the reservoir bottom are more dispersed and therefore, the
curves in the chart are less steep and the transition between water with or without air
bubbles is less sudden than in reservoirs with smaller water depths.
The points in the most left three (green) curves in figure 5.25 marks the points that
the submerged jet collides with the upper part of the submerged pump, see figure 5.26
for an impression. Because the angle of the jet and the fall distance increase with the
water discharge, the location of the plunge point and the bubble plume is not the same
for different discharges. It is possible that a jet with a small discharge does not collide
with the pump whereas a jet with bigger water discharges does.

In practice it is possible there are problems with air bubbles in pressurized pipe
lines with Qap ≥ 0.1 l/min.

Figure 5.26: Impression of a pump air intake discharge experiment. The bubble plume collides
with the submersible pump

75
When looking at the percentages of the entrained air volumes that are taken in by
the pump, the following figures 5.27, 5.28 and 5.29 appear.

Figure 5.27: Percentage of entrained air taken in by the pump with water depth Hw = 0.40 m

Figure 5.28: Percentage of entrained air taken in by the pump with water depth Hw = 0.60 m

76
Figure 5.29: Percentage of entrained air taken in by the pump with water depth Hw = 0.84 m

The curves are often defined by a small number of data points. The first reason is
the relative small number of air entrainment discharge data points between the
measurable air discharges Qa ≤ 100 l/min and the reliable air discharges Qa ≥ 1 l/min,
see figure 5.17. Another cause is the small discharge range between little and much
pump air intake discharges, see figures 5.23 and 5.24. Only data that are in both
figures (5.17 and 5.22 to 5.24) have a data point respectively in figures 5.27 to 5.29.
In some cases where data points have no exact water discharge match in both air
entrainment discharge and pump air discharge measurements, but have close neighbor
discharge data points, an estimation of the entrained air discharge is made to increase
the number of data points in figures 5.27 to 5.29.

With the used uncertain data of pump air intake discharges, the pump air intake
percentages are even more uncertain. Because of the uncertainty and the sensitivity of
the model set-up for these measurements, no clear statements can be made other than;
it is possible to take in up to 30 % of the entrained air by the pump in this model set-
up. Even higher percentages are possible when a pump is used which is more capable
in pumping a gas-liquid mixture. But that subject will not be dealt with in this thesis.

The values of the pump air discharges and the pump air percentages are
conservative when looking at the impeller revolution rate. When the impeller works
for every discharge in its best efficiency point, the results of the pump air discharge
and pump air percentage will be higher, especially at small water discharges. See
figure 5.22.

77
Chapter 6: Conclusions
After doing the experiments with free overfall jets from circular channels, a lot of
interesting phenomena occur that were not expected beforehand.

Literature describes that jet surface disturbances are the main cause of the air
entrainment with plunging jets. These disturbances entrap air at the location between
the jet surface and the reservoir water surface when the surfaces do not fit exactly.
This experimental research found out that the perimeter shape of the intersection
between the falling jet and the reservoir water surface (plunge point) is also an
important factor for the amount of air entrainment. For free overfall jets from circular
channels the horizontal jet cross section shape is at first U-shaped and becomes along
the fall gradually circular. Therefore the perimeter length decreases with increasing jet
fall heights and so the air entrainment discharge decreases too. But then again the jet
surface disturbances grow with the fall height, which enlarges the air entrainment.
The found dimensionless empirical formula for the circular fall height is:
H cir Qw
= 102.0 , see paragraph 5.3.
D g D 2.5

The next surprising results are about the penetration depth of the entrained air
bubble plume. In advance, a constant or increasing penetration depth was expected
with increasing jet fall heights, but it appeared that the opposite effect was more often
the case. The amount of air that is entrained by the plunging jet determines mostly the
penetration depth along with the jet water discharge. The higher the entrainment ratio
air/water discharge is, the smaller the penetration depth. That is because the entrained
air bubbles break up the submerged jet and so the jet loses its momentum. To
illustrate it; when the entrainment ratio is high, the jet lands on a cushion of rising air
bubbles.

The aeration length appeared to be independent of the plunging jet fall height. The
aeration length is mainly affected by the jet water discharge. The empirical maximum
aeration length with an exceedance frequency of once every eight to ten seconds is:
s
La ,max = 0.5Qw0.5 , with the dimension of constant 0.5 equal to
m

From experiments with pump air intake experiments the following observations
are made. Air bubbles are only taken in by the pump when they are in the direct
neighborhood of the pump inlet. The influence sphere of the pump is for discharges
Qw ≤ 15.0 not more than a radius of 0.10 m around the pump inlet centre.

The last unexpected phenomenon is the effect of a pre-rotation at the pump inlet
on the pump air intake discharge. A pre-rotation is a circulation in front of the pump
inlet as a result of the pump impeller revolution rate that is not working in its Best
Efficiency Point (B.E.P.). The pre-rotation pushes away the air bubbles near the pump
inlet and hinders air from being taking in by the pump. But this effect is only marginal
with impeller revolution rates bigger than twice the B.E.P. and is therefore no
effective and economical way to prevent air bubbles being taking in by the pump.

78
In this thesis a couple of mostly empirical formulas in literature are compared
with the results of the experiments. Almost all formulas proposed in literature are only
representative for small jets produced with nozzles. Only Dey (1998, 2001) and
Aigner (1999) have described some recent findings about free overfall flows in
circular channels. But they give no information about the falling jets produced with
the overfall channel flows, nor the air entrainment with the plunging jets.
The only formula from literature that turned out to be valid for the experiments
done is the dimensionless discharge formula of Rajaratnam & Muralidhar (1964):
11
Qw  he  6
= 1.54   . All other formulas are not useful or applicable for this specific
g D 2.5 D
model or the magnitudes of the model dimensions do not correspond to the validity
dimensions of the empirical formulas in literature.

The measured results are obtained with this specific model set-up. Other similar
models can give different results, especially with the pump air intake experiments.
The measurements are very sensitive to all kinds of variations, for example the
turbulence level in the approach flow and the influence of the pump impeller
revolution rate, see figure 5.22.Therefore, the results are only useful for giving an
indication of the in reality occurring orders of magnitudes.

The measured air entrainment discharges Qa differ very much for different falling
jets. The minimum air entrainment is 0.0 l/min and the maximum is at least 300 l/min.
The measured pump air take in discharge is between 0.0 and 10 l/min. With a pump
air discharge more than 10 l/min, the pump stopped working.

The most robust and effective way to prevent air bubbles coming deep into the
sewer sump near the submersible pump inlet is to break up and bend the vertical
submerged jet in the reservoir. When the submerged jet hits a horizontal object below
the water surface, the high velocity flows in the reservoir are bended horizontal. Air
bubbles cannot be brought deeper into the reservoir and rise up unhindered in the
horizontal flows to the water surface. Because real sewer sumps have alternating
water levels, the horizontal object in the reservoir should be below the lowest water
level.

79
Chapter 7: Recommendations
This master thesis is an experimental research about a specific subject which is hardly
described in literature. Because of that, this thesis is mainly a basic description of all
occurring phenomena with the used model set-up. As a result of the numerous
imposed and measured parameters there was insufficient time to examine every
parameter thoroughly. Maybe other research programs can expand the results with
better predicting and more formulas. The experiment program is easily expandable
too. See following list for recommendations for future researches:
• A longer approach flow tube to a minimum of 35 times the hydraulic diameter
of the highest water discharge, to get better and more reliable results.
• An approach flow tube diameter other than D = 0.20 m, to measure the
diameter dependency.
• An inclined approach flow channel, to see the various effects on the air
entrainment.
• Another approach flow channel geometry, to alter the jet cross section shape.
• A deeper reservoir depth, to get more penetration depth results, especially
from jets with small air/water entrainment ratios.
• A wider or bigger reservoir to minimize the reservoir circulation and the
accompanying effect on the air bubble plume.
• A non-axis-symmetrical approach flow suspension compared to the reservoir,
to measure the effects of increased horizontal reservoir circulations, in
particular for pump air intake discharge measurements.
• A disconnected pump water and plunging jet water discharge, to be able to
vary only one parameter. Like real sewer sumps, experiments with non-
stationary reservoir water levels are then possible too.
• Other water properties, like lower water surface tensions and viscosities, to see
the effects on the air entrainment and the air bubble plume.
• A better device or method to measure the plunge point area, to get extra
information and starting-points for an air entrainment formula.

80
References
Aigner, D., Cherubim, C., Overflow in a cylindrical pipe bend. VII-th Conference Problems of Hydro
Engineering . Wroclaw-Szklarska Poreba, Poland, 1999, p.136-141.

Battjes, J.A., Vloeistofmechanica. Lecture notes CT2100 of faculty Civil Engineering and Geosciences
at Delft University of Technogoly, Delft, 2001.

Biń, A.K., ‘Gas entrainment by plunging liquid jets’. Chemical Engineering Science, 48, No.21 (1993),
p.3585-3630.

Burgess, J.M., Molloy, N.A. and McCarthy, M.J., ‘A note on the plunging liquid jet reactor’. Chemical
Engineering Science, 27 (1972), p.442-445.

Calvert, J.B., Open-Channel Flow. July 2003. October 2006.


http://www.du.edu/~jcalvert/tech/fluids/opench.htm

Chanson, H., Hydraulic design of stepped cascades, channels, weirs and spillways. Oxford: Pergamon,
1994.

Chanson, H., The hydraulics of open channel flow: An introduction. 2nd edition. Oxford: Elsevier
Butterworth-Heinemann, 2004.

Detch, R.M., Sharma, R.N., ‘The critical angle for gas bubble entrainment by plunging liquid jets’, The
Chemical Engineering Journal, 44 (1990), p.157-166.

Dey, S., ‘End Depth in Circular Channels’. Journal of Hydraulic Engineering, 124, No.8, (1998),
p.856-863.

Dey, S., ‘EDR in Circular Channels’. Journal of Irrigation and Drainage Engineering, 127, No.2
(2001), p.110-112.

Dey, S., ‘Free overall in open channels: state-of-the-art review’. Flow Measurement and
Instrumentation, 13 (2002), p.247-264.

Flygt, brochure: Top 50, 65, 80, 100, 150, Turnkey pump stations. ITT Industries

Ito, A., Yamagiwa, K., Tajima, K., Yoshida, M. and Ohkawa, A., ‘Maximum penetration depth of air
bubbles entrained by vertical liquid jet’. Journal of Chemical Engineering of Japan, 33, No.6
(2000), p.898-900.

Kusabiraki, D., Murota, M., Ohno, S., Yamagiwa, K., Yasuda, M. and Ohkawa, A., ‘Gas entrainment
rate and flow pattern in a plunging liquid jet aeration system using inclined nozzles’. Journal of
Chemical Engineering of Japan, 23 (1990), p.704-710.

Lubbers, C.L and Clemens, F.H.L.R., ‘Air and gas pockets in sewerage pressure mains’, IWA journal
Water Science and Technology, volume 52, issue 3, p37-44, ISSN Print: 0273-1223 (2005).

Lubbers, C.L and Clemens, F.H.L.R., ‘Breakdown of air pockets in downwardly inclined sewerage
pressure mains’, IWA journal Water Science and Technology, volume 54, issue 11-12, p.233-
240, ISSN Print: 0273-1223, (2006).

McKeogh, E.J., Ervine, D.A., ‘Air entrainment rate and diffusion pattern of plunging liquid jets’.
Chemical Engineering Science, 36 (1981), p.1161-1172.

Nakasone, H., ‘Study of Aeration at Weirs and Cascades’. Journal of Environmental Engineering, Vol.
113, No.1 (1987), p.64-81.

81
Pol-O.R. Bear, Personal Air Conditioning System. November 2006.
http://www.polorbear.net/Coan_eff.html

Sande, E. van der, Smith, J.M., ‘Surface entrainment of air by high velocity water jets’. Chemical
Engineering Science, 28 (1973), p.1161-1168.

Sande, E. van der, Smith, J.M., ‘Mass transfer from plunging water jets’. The Chemical Engineering
Journal, 10 (1975), p.225-233.

Sande, E. van der, Smith, J.M., ‘Jet break-up and air entrainment by low velocity turbulent water jets’.
Chemical Engineering Science, 31 (1976), p.219-224.

Theylaert, A., Rosink, J.G. and Massink, H., Aquatische ecotechnologie, practicum handleiding.
Hogeschool Zeeland, November 2005. Januari 26th 2007.
http://www.hzeeland.nl/~mass0001/bestanden/web_t12_labhandleiding.doc

Yamagiwa, K., Mashima, T., Kadota, S. and Ohkawa, A., ‘Effect of liquid properties on gas
entrainment behavior in a plunging liquid jet aeration system using inclined nozzles’. Journal of
Chemical Engineering of Japan, 26, No.3 (1993), p.333-336

Zhu, Y., Oğuz, H.N. and Prosperetti, A., ‘On the mechanism of air entrainment by liquid jets at a free
surface’. Journal of Fluid Mechanics, 404 (2000), p.151-177.

Other references:
Picture of a V-notch:
http://civcal.media.hku.hk/yuenlong/model_design/instruments/_vnotch.htm

82
Appendix A: Schematic drawing of model set-up

83
Definitions:
• Reservoir with B (m)
• Approach tube diameter D (m)
• Aeration zone La (m)
• Bubble zone Lb (m)
• Fall distance Ljx (m)
• Approach tube length Lp (m)
• Tube–pump distance Ltp (m)
• Critical depth hc (m)
• End depth he (m)
• Fall height H (m)
• Real fall height He (= H + 2/3 he) (m)
• Penetration depth Hp (m)
• Water depth Hw (m)
• Total plunge velocity Vj (m/s)

84

You might also like