The Circular Hydraulic Jump The Influence of Downstream Flow On The Jump Radius

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

The circular hydraulic jump; the influence of

downstream flow on the jump radius


Cite as: Phys. Fluids 34, 072111 (2022); https://doi.org/10.1063/5.0090549
Submitted: 07 March 2022 • Accepted: 24 May 2022 • Published Online: 14 July 2022

Rajesh K. Bhagat and Paul F. Linden

COLLECTIONS

Paper published as part of the special topic on Kitchen Flows

ARTICLES YOU MAY BE INTERESTED IN

Fast dynamics of surfactant probed by the acoustics of a drop impact


Physics of Fluids 34, 072107 (2022); https://doi.org/10.1063/5.0098642

Numerical simulations of cell sorting through inertial microfluidics


Physics of Fluids 34, 072009 (2022); https://doi.org/10.1063/5.0096543

Viscous Rankine vortices


Physics of Fluids 34, 073603 (2022); https://doi.org/10.1063/5.0090143

Phys. Fluids 34, 072111 (2022); https://doi.org/10.1063/5.0090549 34, 072111

© 2022 Author(s).
Physics of Fluids ARTICLE scitation.org/journal/phf

The circular hydraulic jump; the influence


of downstream flow on the jump radius
Cite as: Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549
Submitted: 7 March 2022 . Accepted: 24 May 2022 .
Published Online: 14 July 2022

Rajesh K. Bhagata) and Paul F. Linden

AFFILIATIONS
Department of Applied Mathematics and Theoretical Physics, Wilberforce Road, Cambridge CB3 0WA, United Kingdom

Note: This paper is part of the special topic, Kitchen Flows.


a)
Author to whom correspondence should be addressed: rkb29@cam.ac.uk

ABSTRACT
In this study, we examine the consistency of a gravity-based predictive theory for a hydraulic jump, given by Kurihara [Proceedings of the Report
of the Research Institute for Fluid Engineering (Kyusyu Imperial University, 1946), Vol. 3, pp. 11–33]; Tani [J. Phys. Soc. Jpn. 4, 212–215 (1949)]
with the phenomenological condition at the jump given by Rayleigh [Proc. R. Soc. London, Ser. A 90, 324–328 (1914)]; and Watson [J. Fluid
Mech. 20, 481–499 (1964)] and show that in light of experimental evidence, the gravity-based predictive theory for the kitchen sink hydraulic
jump is incompatible with the phenomenological condition, which must be valid. We also examine the solution to the downstream film and its
potential influence on the hydraulic jump. We show that for all practical purposes, at normal flow conditions, the downstream liquid film remains
flat and does not affect the jump, and the theory given by Bhagat et al. [J. Fluid Mech. 851, R5 (2018)] gives an excellent prediction of the jump
radius. For high viscosity liquids, on a relatively large plate, the viscous dissipation in the downstream film could increase the jump height and,
consequently, move the jump radius inward.
Published under an exclusive license by AIP Publishing. https://doi.org/10.1063/5.0090549

I. INTRODUCTION singular at a finite radius, where the wave speed equals the flow speed
The circular hydraulic jump is a common phenomenon observed U. This condition
pffiffiffiffiffi was expressed in terms of the local Froude number
when we turn on the tap of a kitchen sink; the water from the tap falls Fr  U= gh ¼ 1, where g is the acceleration due to gravity and h is
as a vertical jet onto the bottom of the sink and spreads radially in a the fluid depth. However, when compared with experiments carried
thin film. At a certain distance from the point of impact, typically a out using water, the theory3,4 significantly over-predicted the jump
few centimeters, the film thickness increases abruptly, forming what is radius.5
conventionally known as a hydraulic jump (see Fig. 1). These jumps In the case of the thin-film jump observed on the scale of the
show visual resemblance to the hydraulic jumps seen in rivers and kitchen sink, surface tension produces a force on the spreading film.
bores, and an early mathematical model in 1914, aimed at describing Bush and Aristoff6 extended the earlier model to include the influence
bores, was built on the concept of momentum conservation,1 and the of surface tension and applied it to tabletop scale laboratory experi-
circular hydraulic jump was classified as a bore at small scale. From ments, where the jet impact occurred in a reservoir initially filled with
this viewpoint, the hydraulic jump is regarded as a sharp transition in water. On impact of the jet, the liquid spilled over a weir at the reser-
water depth—a standing wave and the stationary counterpart of a tidal voir boundary that determined the height of the outer film. For a given
bore—and, at the jump, the phenomenological condition, in which the jet flow rate, increasing the height of the weir produced a larger
rate of change of momentum in the flowing stream should be balanced gravitational pressure, which moved the jump radius inwards, and at a
by the thrust of pressure produced by order of magnitude increase in certain height, the jump entirely disappeared. This experiment in con-
the depth of the water, was satisfied.2 The first predictive model was junction with the phenomenological model gives an impression that,
developed in the 1940s by Tani3 following Kurihara.4 They proposed even at the scale of a tabletop, gravity is the principal force and the key
that a hydraulic jump is caused by flow separation due to an adverse to the jump dynamics.
gravitational pressure gradient. They modeled the flow in terms of the In addition, in thin-film flows, viscosity can decelerate the fluid
shallow water equation and showed that the equation becomes and, by volume conservation, increase the film depth producing a

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-1


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

II. THEORY
A. Consistency and compatibility between the
phenomenological and the predictive models
The model based on the momentum theorem given by Rayleigh1
and Watson2 requires the momentum flux to be balanced by pressure
difference across the jump
1
u2 h  U 2 H ¼ gðH 2  h2 Þ; (1)
2
where u, h and U, H are the speed and depth upstream and down-
FIG. 1. Schematic depicting a cross section of a circularly symmetric hydraulic stream of the jump, respectively, the overline represents a depth aver-
jump and its key features; radially spreading super and subcritical films conjoined age, and we have taken the fluid density q ¼ 1 for convenience.
by a sharp jump. At the jump in the supercritical film, the Froude number is defined Ignoring the influence of surface tension, the gravity-based pre-
as Fr, whereas in the subcritical film, it is defined as Frex. dictive model given by Tani3 and Kurihara,4 which is based on the the-
ory that separation of the flow from the surface due to the adverse
further adverse pressure gradient. Recently, Wang and Khayat7 gravitational pressure gradient, produces the jump, gives for the radial
rewrote the equation given by Tani3 and Kurihara4 in parametric form gradient of the film thickness
and concluded that the model is capable of predicting jumps in high  
viscosity liquids.5 Q 0 1 Q 2 1
Recently, however, experimental evidence has cast doubt on the  f ð0Þ 2  C2
dh 2pC1 rh 2pC1 r 3 h
role of gravity in these thin-film jumps. Experiments on the normal ¼  2
dr Q 1
impact of a jet onto a planar surface showed that the radius of the C2  gh
jump was independent of the orientation of the surface to gravity.8 2pC1 r 2 h2
 
Bhagat et al.5 provided further experimental evidence, including  Q 0 1 C2 Q 2 1
f ð0Þ 2 
experiments carried out in micro-gravity on-board a “zero-gravity” gh 2pC1 rh gh 2pC1 r 3 h
flight and in a drop tower, showing the formation of jumps in micro- ¼ 2
; (2)
ðFr  1Þ
gravity occur at the same location as in the terrestrial experiments. It is Ðh Ðh
impossible to reconcile these observations with the notion that gravity where u  ¼ C1 us ¼ h1 0 udz; u2 ¼ C2 u2s ¼ h1 0 u2 dz, us is the surface
qffiffiffiffi
plays a central role in the formation of these thin-film jumps. 2
velocity, and Fr ¼ ugh ; f 0 ð0Þ arises from the velocity profile ansatz
Nevertheless, the idea that gravity is central to the formation
of these jumps persists. Bohr et al.12 developed a theory based on where sw ¼ l uhs f 0 ð0Þ is the wall shear stress.
the shallow water equations, which essentially recovers the theory Here, we examine the consequences of applying (2) to the
formulated by Tani,3 Kurihara,4 and their key result—the scaling momentum conservation (1), noting that we expect a significant
relationship for the jump radius, R  Q5=8  3=8 g 1=8 —can be increase in the flow depth after the jump (i.e., h  H). We consider
retrieved by imposing the jump condition, Fr ¼ 1.13 Based on a pos- two approximate but reasonable flow profiles. The first is laminar flow
tulation that, the exit Froude number, Frex, at the jump, in the with the same velocity profile before and after the jump, and the sec-
downstream film remains constant, Duchesne et al.14 obtained a ond is turbulent flow after the jump, resulting from flow separation,
modified version of the scaling relationship given by Bohr et al.,12 with uniform velocity across the depth.
and apparently confirming experiments have also been reported.12 Laminar flow: Assuming that the shape factor is the same on
So what is the issue here? Measurements of the jump radius are not either side of the jump, applying continuity of volume flux yields
unequivocal. There are experimental errors, even though the jump
radius is a simple measurement. More challenging is the reconcilia-  Rh ¼ U
u  RH ¼ Q
2p (3)
tion of the use of fluids with different surface tensions and viscosi- ) C1 us h ¼ C1 Us H:
ties. Conceptually, the most difficult problem is the role of the flow qffiffiffiffi
downstream of the jump. As discussed above, the difference in fluid u2
Substituting the condition of criticality given in (2), Fr ¼ gh ¼ 1,
depths before and after the jump provides a pressure difference that
and (3) into (1) yields
must be matched by the change in momentum of the flow.
Experiments are carried out on a plane of finite dimensions—what 1
u2 h  U 2 H ¼ gðH 2  h2 Þ: (4)
role, if any does that play? 2
In this paper, we discuss the different theories and experimental After some algebra, we find
evidence that have led to this somewhat confused picture and attempt   
h h
to reconcile them and to clarify the dynamics. In Sec. II, we examine  1 2 þ 1 ¼ 0: (5)
the early theories and their relation to the momentum conservation H H
constraint in order to determine the role of the downstream conditions Equation (5) gives Hh ¼ 1, which is a trivial solution of (1) and there is
on the jump formation. In Sec. III, we discuss the experiments and no jump.
develop a consistent theory for all cases. Finally, in Sec. V, we present Turbulent flow: when the appropriate velocity profile across the
our conclusions. jump is a turbulent profile,2 we write

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-2


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

 Rh ¼ URH
u HðrÞ2 dHðrÞ
(6) UðrÞ ’  g ; (13)
) C1 us h ¼ UH; 3 dr
Q
and (5) could be written as and combining (13) with volume flux conservation UrH ¼ 2p yields
1
u2 h  U 2 H ¼ ; (7) dHðrÞ
’
3Q
: (14)
2 dr 2pgrH 3
2 3 2
2C1 h h
 3 2 þ 1 ¼ 0; (8)
C2 H 3 H Solving (14) for the boundary condition H ¼ H1 for r ¼ R1 gives
 !14
3 2
h h
1:53 3  3 2 þ 1 ¼ 0: (9) Q R1
H H 4
HðrÞ ¼ H1 þ 6 ln : (15)
This equation yields three real roots, Hh ¼ 0:51; 0:76; 1:75, and so pg r
the only reasonable solution is h ¼ 0:76H. It is instructive to compare (14) and (10), the two mathematical formu-
In summary, this analysis shows that in order for inverse gravita- lations describing the downstream film. Note that balancing frictional
tional pressure gradient—that could cause the flow separation and loss with gravitational pressure ignores the terms arising from circular
consequently the jump—to build, the depth of a supercritical film has expansion—the first term in (10), ð2pC Q 2 1
Þ gr3 h2 —and consequently, in
to be of the same order as that of the subcritical film, which is not 1

experimentally observed. contrast to (10) where the slope, dH


dr , could either be positive or negative,
(14) gives a negative slope, that we will show, is not observed in experi-
R1
B. Viscous effects ments. They further assumed that H1 4
 6 Qpg ln r so that

Here, we examine the effect of viscous dissipation on the height  !14


Q R1
of the jump and, consequently, the jump radius. Again consider the HðrÞ ’ 6 ln : (16)
gravity based model (2). In the flow downstream of the jump, pg r
Fr 2 ! 0, and in this limit, Combining (12) and (16) gives for the dependence of the jump radius
  
dh Q Q 1 0 1 on the plate radius
¼  f ð0Þ : (10)
dr 2pC1 grh2 2pC1 r 2 h   3=8
R1 1 5=8 3=8 1=8
From (10), dh R ln ¼ bFrex Q  g : (17)
dr ! 0, at the edge of a large plate r ! 1, and for R
r 2
2f 0 ð0ÞpC1 HQ > 1 the slope of the film is negative and vice versa. Combining the constant exit Froude number with momentum conser-
Thus, we expect the slope of the surface downstream of the jump to be vation across the jump, it can be shown that the supercritical thin film
negative and to increase in magnitude as the viscosity increases. Froude number at the jump also remains constant. For the laminar
Furthermore, (10) allows us to investigate the slope of the film velocity profile on either side of the jump, Fr ¼ 4.18, and for a laminar
near the jump, where it is expected to be the most pronounced. For supercritical film and a turbulent subcritical film, Fr ¼ 7.11.
typical experimental flow rates Qðoð105 m3 s1 ÞÞ for which In summary, the constant exit Froude number hypothesis14 gave
R2
r ¼ Rj ðoð102 mÞÞ, and h ¼ Hj ðoð103 mÞÞ, yielding 2f 0 ð0ÞpC1 Hj Qj a number of predictions that can be examined in order to test the
validity of the hypothesis and the theory, namely, R / g 1=2QH 3=2 ; H
 ðoð  104 ÞÞ, and so for a liquid with kinematic viscosity
R1 14
ðoð106 m2 s1 ÞÞ (for example, water), near the jump, the slope ’ ð6 Q
pg ln r Þ and the jump Froude number is around 5.
remains positive and tends asymptotically to zero, while for a liquid
with viscosity  > 104 m2 s1 , could have a negative slope that could D. A unified theory
asymptotically reach to zero.
A theory that includes the effects of gravity, surface tension, and
viscosity was presented by Bhagat et al.8 We briefly summarize it here
C. Effect of plate size
but do not give details. Combining Eqs. (5.1) and (5.6) of Bhagat
In order to examine the effect of a finite plate, Duchesne et al.14 et al.,8 (5.6) can be reformulated to obtain the relation for the slope of
formulated another theory, which also gave good agreement with their the surface, given by
experimental data. The empirical evidence obtained from their experi-  
ments prompted them to postulate that the exit Froude number in the Q 1 Q 2 1
lf 0 ð0Þ  C 2 q
downstream film at the jump (see Fig. 1) is Frex ¼ 0:33 6 0:01, dh 2pC1 rh2 2pC1 r 3 h
¼  2
U Q dr Q 1 c 1
Frex ¼ pffiffiffiffiffiffi ¼ ¼ 0:33 6 0:01; (11) C2 q   qgh
gH 2pRg 1=2 H 3=2 2pC1 r 2 h2 h 2
 
which can be re-arranged to give an expression for the jump radius R, f 0 ð0Þ 2pC1 h
r
Q C2 Q r
R¼ : (12)  ; (18)
Frex 2pg 1=2 H 3=2 1 1
1  2
Assuming a parabolic velocity profile and balancing gravitational We Fr
pressure and viscous friction, they also formulated a simplified equa- where viscous effects are included via the wall stress sw and we define
tion for the downstream liquid film thickness the Weber number and the Froude number as

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-3


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

sffiffiffiffiffiffiffiffiffiffiffiffi
C2 qu2s h 2C2 u2s
We  ; Fr  ; (19)
c gh

respectively.
Consequently, from (18), the flow is singular, and the jump
occurs when
1 1
þ ¼ 1: (20)
We Fr 2
Furthermore, Bhagat et al.8 have shown that for water and other low
viscosity and high surface tension liquids, in the supercritical film at
the jump, Fr 2  1 and the jump occurs when We  1. Bhagat et al.5
have shown that combining (1) and the jump condition, We  1 gives
the height of the jump
sffiffiffiffiffi
2c
H : (21)
qg

Note that in the absence of gravity, the jump condition is We ¼ 1,


while in the absence of surface tension it is Fr ¼ 1, as found before.3 FIG. 2. The liquid film thickness obtained by solving Eqs. (2) and (18), for a water
jet at a flow rate Q ¼ 2 l min1 . For a subcritical film, Eqs. (2) and (18) give an
Consequently, this unified theory of Bhagat et al.8 provides the jump almost identical solution, and the liquid film remains flat. The vertical line demar-
condition that includes surface tension, gravity, and viscosity. In Sec. cates the experimental location of the jump. While (2) over-predicts the jump, (18)
III, we discuss the experimental measurements in this context. gives an excellent prediction.
III. RESULTS
Refs. 12 and 15 for three different flow rates. Again excellent agree-
We now examine data from published experiments (see Table I)
ment for the jump radius is observed.
and connect them to the theory presented above. We examine the
A further comparison of the models and observations obtained
jump radius, the depth of the flow downstream of the jump and its
by Mohajer and Li16 is shown in Fig. 4. We see, as before that (2)
dependence on the height of the weir and the size of the plate, and the
which ignores the influence of surface tension overpredicts the jump
constant Froude number hypothesis. radius while the unified theory (18) gives excellent agreement. Both
(2) and (18) give almost identical predictions for the downstream film
A. Experiments with water
Figure 2 compares the measured film thickness produced by ver-
tical impingement of a water jet falling on to a horizontal surface with
(2) and (18). In the figure, the vertical line demarcates the experimen-
tally measured location of the hydraulic jump. In the supercritical
region, both models agree well with the measured thickness of the liq-
uid film but only (18) gives an accurate prediction of the jump radius,
while (2) predicts almost twice the radius. In the subcritical region, (2)
and (18) are almost identical, while (21) gives excellent prediction of
the observed depth. It is worth noting in this case of relatively low
viscosity, the downstream film thickness is almost constant.
In Fig. 3, we show effect of changing the jet flow rate. The liquid
film profiles, given in (18) for both the super and subcritical regions
are plotted and compared with the observed jump radii obtained from

TABLE I. Experimental data used for comparison.

Q c  H
Reference Liquid (cm3 s1 ) (cm3 s1 ) (m) (m)

16 Water 2.5–8.33 3058 3.83 3.92


16 Water and surfactant 1.66–3.33 807.4 2.75 2.75 FIG. 3. The effect of flow rate. The solution of (18) is plotted for three different flow
rates and compared against the experimentally measured values of hydraulic jumps
17 Ethylene glycol solution 27 129.7 2.86 2.76 obtained from Refs. 12 and 15. The vertical lines demarcate the location of experi-
14 Silicone oil-1 4.3 12.44 2.37 2.1 mentally measured jumps, which is in excellent agreement with the theory. The
downstream film remains flat.

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-4


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

thickness, and (21) gives excellent prediction to the experimentally


measured value of the depth of the film.

B. Experiments with other liquids


Bohr et al.17 measured the liquid film thickness for impingement
of the jet of an ethylene glycol solution, which has about half the sur-
face tension and seven times the viscosity of water. We find in Fig. 5
again that (2) overpredicts the jump radius, while (18) gives excellent
predictions of both the jump and the supercritical film thickness and
(21) gives H ¼ 2:76 mm in good agreement with observations.
The effect of viscosity is shown by experiments by Duchesne
et al.,14 who measured the liquid film thickness and hydraulic jump
produced by a jet of silicon oil with a viscosity of 20–100 times that of
water. In this case, the experimental data (Fig. 6) show that the viscous
dissipation in the downstream film plays a significant role, and, due to
the increased jump height, the jump moved slightly inwards, and both
theories slightly over-predict the jump; nevertheless, (18) gives better
prediction. For these low surface tension and high viscosity liquids, the
surface energy contribution remains relatively low while the decelerat- FIG. 5. The figure depicting comparison between the experimental and the pre-
dicted film thickness for an ethylene glycol solution jet at flow rate
ing liquid film becomes thicker, and consequently, gravity could play a Q ¼ 27  106 m3 s1 , density q ¼ 1100 kg m3, kinematic viscosity density
significant role.  ¼ 7:6  106 m2 s1 , and surface tension, c ¼ 45 mN m1 . The experimental
data are obtained from Ref. 17. While both Eqs. (2) and (18) give good prediction
C. Effect of a weir to liquid film thickness, (2) fails to predict the jump. Equation (18) gives excellent
prediction to film thickness as well as the jump.
So far, we have compared experimental jump radii with the theo-
ries; Bohr et al.17 measured the liquid film thicknesses produced by
impingement of an ethylene glycol solution jet with different outer film outer film boundary condition, the films are approximately flat.
thickness boundary conditions, and showed that beyond jump, the Figure 8 compares the outer film thickness obtained for a flow rate
height of the liquid film remained almost constant (see Fig. 3 of Ref. 17). Q ¼ 1:62 l min1 and different outer boundary conditions for the
In Fig. 7, we present the computed outer film thickness for water outer film height. Except for the minimum film thickness for which
qffiffiffiffi
on a relatively large plate (the diameter of the plate ¼ 1 m) and a the boundary condition is taken to be qg2c
¼ 2:76 mm, even on a
flow rate Q ¼ 2 l min1 , and one could see that irrespective of the

FIG. 6. The figure compares predicted and the experimental liquid film thickness
FIG. 4. The figure depicting comparison between the experimental and the pre- produced by impingement of a silicon oil jet of density q ¼ 950 kg m3, kinematic
dicted jump location and the downstream film thickness for water at a flow rate of viscosity density  ¼ 20:4 6 0:6  106 m2 s1, and surface tension, c
Q ¼ 3:75  106 m3 s1 . The experimental data are obtained from Ref. 16. While ¼ 20 mN m1 . The experimental data are obtained from Ref. 14, at a flow rate
both Eqs. (2) and (18) give good prediction to the liquid film thickness; (2) fails to Q ¼ 17  106 m3 s1 . While both Eqs. (2) and (18) give good prediction to
predict the jump. (18) gives excellent prediction to film thickness as well as the liquid film thickness, (18) gives a better prediction of the jump, despite the signifi-
jump. cant influence of viscous dissipation in the downstream film.

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-5


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 7. The effect of boundary conditions. Hydraulic jump experiments are often FIG. 9. The effect of viscosity. For low viscosity liquids, Eqs. (2) and (18) give
conducted in reservoirs where the height of outer film can be varied. We compare approximately an identical solution for the downstream films. For high viscosity
the predicted downstream film profiles for water (Q ¼ 2 l min1 ) with three different liquids, (18), accounting for surface as well as gravitational energy give a higher
boundary conditions, H1 ¼ 4; 6; and 8 mm, respectively. The liquid film profile film thickness compared to (2).
remains unaffected by the change in the boundary condition.
large plate, the viscous dissipation in the downstream film is likely to
relatively large plate, the liquid films remain flat. Furthermore, the sol- move the hydraulic jump inwards. Indeed, Bhagat et al.5 found that
utions obtained from (2) and (18) superimpose each other. the measured jump radii for liquid with high viscosities were slightly
Figure 9 compares the liquid film thickness obtained for a liquid smaller than the predicted values. We would like to reemphasise that,
of surface tension c ¼ 72 mN m1 and the kinematic viscosity range in order to examine the effect of viscous dissipation, we have chosen a
between 106 and 103 m2 s1 . For liquids with low kinematic viscosi- rather large plate, which for most practical scenarios, even for high vis-
ties, in this scenario, Eqs. (2) and (18) give an identical solution while cosity liquids, on smaller plates viscous dissipation remains
for liquid with high kinematic viscosities, the solution differs; (18) insignificant.
including surface tension gives slightly higher values compared to (2). Figure 10 examines the effect of surface tension; increasing the
In the case of a liquid with high kinematic viscosity on a relatively surface tension gives a slightly higher value for the liquid film thick-
ness but the effect of surface tension is minimal. In Ref. 13, we have
described that for thick films, the rate of change of surface energy is
low, and for flat films, it is zero, which concurs with our results here.

FIG. 8. The effect of boundary conditions. We compare the predicted downstream


film profiles for an ethylene glycol solution jet with flow rate Q ¼ 2 l min1 , density
q ¼ 1100 kg m3 , kinematic viscosity density  ¼ 7:6  106 m2 s1, and surface
tension, c ¼ 45 mN m1 . In order to pronounce the viscous dissipation effect, we FIG. 10. The effect of surface tension. Increasing surface tension for a high viscos-
consider a plate of diameter 1m and four different boundary conditions, H1 ity liquid gives a slightly higher value for the downstream film, but the effect is mini-
¼ 2:76; 4; 6; and 8 mm, respectively. For a thin film [H(boundary) ¼ 2:76 mm], the mal. For low viscosity liquids, surface tension does not play much significant role,
viscous dissipation does change the jump height. The liquid film profile remains and the two theories give an almost identical solution for the downstream liquid film
unaffected by the change in the boundary condition. thickness.

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-6


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

D. Predictions of the constant Froude


number hypothesis
In Sec. II C, we identified the testable predictions of the constant
Froude number hypothesis given by Duchesne et al.14 that we examine
here. Figure 11 compares the downstream film thicknesses given in
(2), (15), (18), and (16) for water at a flow rate Q ¼ 2 l min1 on a tar-
get plate of diameter 1 m. Equations (2), (15), and (18) give an approx-
imate constant liquid film thickness, except at smaller radii where they
appear to be diverging. The mismatch between (15) in contrast to (2)
and (18) is expected due to the assumption that frictional forces are
balanced by gravity through the free surface slope, (14). Equation (16),
which assumes that the depth of the film at infinity is zero, also fails to
give a reasonable prediction for the liquid film thickness and the jump
radius.
It is worthwhile to note that the jump relationship (17) relies
upon a reasonable prediction of the jump height that Eq. (16) fails to
give and, consequently, for water (17) does not predict the jump
radius, and the constant exit Froude number hypothesis does not
hold. Note that Duchesne et al.14 did not compare their theory with FIG. 12. The comparison between liquid downstream film thicknesses for a silicone
the experiments carried out using water or other low viscosity liquids. oil Silicon oil jet of density q ¼ 950 kg m3, kinematic viscosity density  ¼ 20:4
The theory of Duchesne et al.14 also predicted that the jump 6 0:6  106 m2 s1 , surface tension, c ¼ 20 mN m1 , and at a flow rate Q
radius R / g 1=2QH 3=2 ; in the case of water and other low viscosity liquids, ¼ 2 l min1 on a target plate of diameter 1 m.
the downstream film and jump remain approximately constant that
implies R / Q. However, on existing experimental data, Bhagat et al.5 different geometries and sizes, and for water, the jump radii were
and Bhagat et al.8 clearly show that R / 6 Q. Although for a limited found to independent of the plate geometry and sizes. Finally, the
range of experimental data—for Q; 0  9 cm3 s1 —Mohajer and Li16 constant exit Froude number hypothesis—Frex ¼ 0.33—also implies
did suggest that R / Q but this could be refuted by comparing the that the supercritical film Froude number should be 4.18 or 7.11,
experimental data, in the same range, from other independent research which are arbitrary numbers, and there is no physical basis for this.
groups, including Refs. 8, 12, 15, 18, and 20. Furthermore, a nonlinear Figure 12 compares the height of the downstream film for high
relationship between Q and R also implies that the exit Froude number viscosity and low surface tension silicon oil; in this case, (16) gives a
is not constant. reasonable prediction of the jump height. It is instructive to note that
The theory of Duchesne et al.14 also implied that the jump radius Duchesne et al.14 compared their theory with only high viscosity
is a function of the plate diameter, R ¼ f ðR1 Þ. However, in Ref. 5, we liquids and showed good agreement. In Ref. 5, we have delineated why
compared experimental data for the experiments carried out by inde- the gravity-based theory could give a reasonable prediction for high
pendent research groups over more than three decades on plates of viscosity and low surface tension liquids.
IV. DISCUSSION
It is worthwhile to consider some recent works apparently
explaining why the surface tension-based theory of Bhagat et al.8 was
able to predict the jump. Duchesne and Limat9 argued that in the sce-
nario when the outer film is still developing, the depth of the film is
dependent upon the contact line force arising due to surface tension
and the wetting properties of the substrate, and so in combination
with the phenomenological model, it can be shown that the jump
depends upon surface tension. Essentially, the argument in the model9
suggests that the contact line force could influence the jump. About a
decade ago, Wilson et al.10 gave virtually the same mathematical rela-
tion balancing momentum and surface forces at the jump radius on a
vertical plate, but the model was incapable of explaining why these
jumps are independent of the surface property and the contact angle.
Furthermore, in Ref. 5, we have given experimental evidence showing
why these explanations are untenable; using an experiment, we will
elucidate it further. Avedisian and Zhao11 carried out zero-gravity
experiments in a drop tower, and their experimental setup ensured
FIG. 11. The comparison between liquid downstream film thickness for water at complete wetting and no scope for the emergence of any contact lines.
flow rate Q ¼ 2 l min1 on a target plate of diameter 1 m. For some of their experiments, they impinged liquid jets on deep pools

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-7


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

phenomenological condition—which must be valid—and the predic-


tive theory suggests that for the adverse pressure gradient to be signifi-
cant enough to produce separation, the subcritical liquid film
thickness at hydraulic jumps should be of the order of jump height,
which is not experimentally observed.
We eliminated velocity and reformulated the theory of Bhagat
et al.8 in the form of the film thickness, h and the radial coordinate, r
and compared it with the theory of Tani3 and Kurihara.4 While the
theory of Tani3 and Kurihara4 theory fails to predict the jump radius,
the theory of Bhagat et al.8 gives an excellent prediction for the jump
location as well as the liquid film thickness (see Figs. 2, 3, and 5).
Examining the effect of viscous dissipation in downstream films,
we showed that for low viscosity liquids such as water, for all reason-
able flow rates and plate dimensions, the liquid film after the jump
remains flat, and both the terms arising from viscous dissipation and
expansion of the film eliminate each other, giving a flat film. We also
showed that in the case of high viscosity liquids, viscous dissipation in
the film may become significant, subject to the dimension of the plate.
As a closing remark, we would like to invoke the results recently given
by Bhagat et al.,5 where they compare their theory with more than
three decades of experimental data; they showed that the theory of
Bhagat et al.8 gives excellent prediction to the experimental data for
low surface tension liquids, while slightly over-predicting the high vis-
cosity liquids, showing consistency. Note that these separate experi-
ments were carried out on various plates of unknown dimensions;
nevertheless, they did not affect the jump radius. Finally, we also
examined the theory based on the constant exit Froude number Frex
hypothesis given by Duchesne et al.14 and showed that the theory does
not predict the jump radius for water and other low viscosity liquids,
and Frex is not constant for such liquids.

ACKNOWLEDGMENTS
The authors would like to thank the Leverhulme Trust and the
FIG. 13. The illustrative sketch of a micro-gravity drop tower experiment carried out Issac Newton Trust (No. ECF-2021–196) and Darwin College,
by Avedisian and Zhao11 (see Fig. 11 of their reference). The experiments were University of Cambridge for financial support.
carried out in deep pools of liquid (10 and 15 mm deep). In the terrestrial setting,
the jumps did not appear; however, during the free fall, the jump appeared, irre-
spective of the depth of the pool it appeared at the same location and Eq. (18) AUTHOR DECLARATIONS
gave excellent prediction to the jump (see Ref. 5).
Conflict of Interest
The authors have no conflicts to disclose.
of liquid (see the illustrative sketch, Fig. 13, inspired from Fig. 11 of
Ref. 11), where in normal gravity, the jump formation was suppressed DATA AVAILABILITY
by the large gravitational pressure; however, during free fall, the jump The data that support the findings of this study are available
appeared at a location where local We ¼ 1, giving excellent agreement within the article.
with (18), the theory of Bhagat et al.8 We wish to reemphasise that in
the scenario of an unobstructed flow on a flat plate in terrestrial REFERENCES
experiments, and in micro-gravity experiments, for a given flow rate 1
Lord Rayleigh, “On the theory of long waves and bores,” Proc. R. Soc. London,
Q, the jump remained at the same location. Ser. A 90, 324–328 (1914).
2
V. CONCLUSIONS E. J. Watson, “The radial spread of a liquid jet over a horizontal plane,” J. Fluid
Mech. 20, 481–499 (1964).
In this paper, we examined the effect of downstream flows on the 3
I. Tani, “Water jump in the boundary layer,” J. Phys. Soc. Jpn. 4, 212–215 (1949).
4
hydraulic jumps theoretically and compared it with experiments. We M. Kurihara, “On hydraulic jumps,” in Proceedings of the Report of the
examined the consistency between the phenomenological condition at Research Institute for Fluid Engineering (Kyusyu Imperial University, 1946),
Vol. 3, pp. 11–33.
the jump, given by Watson,2 Taylor19 with Tani,3 and the gravity- 5
R. K. Bhagat, D. I. Wilson, and P. Linden, “Experimental evidence for surface
based theory of Kurihara,4 which postulates that the jump occurs due tension origin of the circular hydraulic jump,” arXiv:2010.04107 (2020).
to adverse pressure gradient produced due to viscous dissipation caus- 6
J. W. M. Bush and J. M. Aristoff, “The influence of surface tension on the cir-
ing the liquid film to become thicker. Combining the cular hydraulic jump,” J. Fluid Mech. 489, 229–238 (2003).

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-8


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

7 14
Y. Wang and R. E. Khayat, “The role of gravity in the prediction of the circular A. Duchesne, L. Lebon, and L. Limat, “Constant Froude number in a circular
hydraulic jump radius for high-viscosity liquids,” J. Fluid Mech. 862, 128–161 hydraulic jump and its implication on the jump radius selection,” Europhys.
(2019). Lett. 107, 54002 (2014).
8 15
R. K. Bhagat, N. K. Jha, P. F. Linden, and D. I. Wilson, “On the origin of Y. Brechet and Z. Neda, “On the circular hydraulic jump,” Am. J. Phys. 67,
the circular hydraulic jump in a thin liquid film,” J. Fluid Mech. 851, R5 723–731 (1999).
16
(2018). B. Mohajer and R. Li, “Circular hydraulic jump on finite surfaces with capillary
9
A. Duchesne and L. Limat, “Circular hydraulic jumps: Where does surface ten- limit,” Phys. Fluids 27, 117102 (2015).
17
sion matter?,” J. Fluid Mech. 937, R2 (2022). T. Bohr, C. Ellegaard, A. E. Hansen, and A. Haaning, “Hydraulic jumps, flow sep-
10
D. Wilson, B. Le, H. Dao, K. Lai, K. Morison, and J. Davidson, “Surface flow aration and wave breaking: An experimental study,” Physica B 228, 1–10 (1996).
18
and drainage films created by horizontal impinging liquid jets,” Chem. Eng. K. Choo and S. J. Kim, “The influence of nozzle diameter on the circular
Sci. 68, 449–460 (2012). hydraulic jump of liquid jet impingement,” Exp. Therm. Fluid Sci. 72, 12–17
11
C. Avedisian and Z. Zhao, “The circular hydraulic jump in low gravity,” Proc. (2016).
19
R. Soc. London, Ser. A 456, 2127–2151 (2000). G. I. Taylor, “The dynamics of thin sheets of fluid. III. Disintegration of fluid
12
T. Bohr, P. Dimon, and V. Putkaradze, “Shallow-water approach to the circular sheets,” Proc. R. Soc. London, Ser. A 253, 313–321 (1959).
20
hydraulic jump,” J. Fluid Mech. 254, 635–648 (1993). J. Stevens and B. W. Webb, “Measurements of flow structure in the radial layer
13
R. K. Bhagat and P. F. Linden, “The circular capillary jump,” J. Fluid Mech. of impinging free-surface liquid jets,” Int. J. Heat Mass Transfer 36, 3751–3758
896, A25 (2020). (1993).

Phys. Fluids 34, 072111 (2022); doi: 10.1063/5.0090549 34, 072111-9


Published under an exclusive license by AIP Publishing

You might also like