You are on page 1of 8

Sustainable Hydraulics in the Era of Global Change – Erpicum et al. (Eds.

)
© 2016 Taylor & Francis Group, London, ISBN 978-1-138-02977-4

Relation between free surface profiles and pressure profiles with respective
fluctuations in hydraulic jumps

J.D. Nóbrega & H.E. Schulz


Department of Hydraulics and Sanitary Engineering, School of Engineering at São Carlos,
University of São Paulo, Brazil

M.G. Marques
Institute of Hydraulic Research, Federal University of Rio Grande do Sul, Brazil

ABSTRACT: Because the hydraulic jump is a dissipative singularity, it is used to dissipate energy for example
in stilling basins downstream of spillways. On the other hand, concerns about stilling basins are the possibility of
cavitation and uplift of baffle blocks, which are mainly related to pressure fluctuations, and the water depths along
the hydraulic jump, relevant for the design of the walls of the stilling basins. Interestingly, pressure fluctuations
and depth fluctuations are caused by the same turbulence in the roller. So, both phenomena may present similar
behaviors for the statistical properties. This paper analyzes instantaneous water depth data, measured using
ultrasonic sensors, for inflow Froude numbers from 2.8 to 5.3. Statistical parameters of the water depth data
were obtained and compared to data of pressure fluctuations found in the literature. As expected, it is shown that
there are similarities between the behavior of both phenomena.

1 INTRODUCTION

Hydraulic jump (HJ) is the open flow phenomenon


related to the transition of a supercritical to a sub-
critical flow condition. The hydraulic jump may be
a useful phenomenon, for example in stilling basins
downstream of dam spillways, where the presence of Figure 1. Hydraulic jump. ∗ Lr = roller length; Ltr =
baffle and chute blocks imposes the hydraulic jump transition length; and Lj = hydraulic jump length.
formation and intense local energy dissipation.
The hydraulic jump is usually described as a com-
plex phenomenon, being characterized by a rapid recirculation pattern (Lr – roller length), with the water
expansion of the free surface level. At the region of at the vicinity of the free surface tending to move down
rapid expansion, a high amount of air enters into the in the opposite direction of the main flow. Beneath
flow in the form of bubbles. The air bubbles are then the roller region, the main flow stream expands in the
advected and diffused in the flow, and finally released downstream direction (see Fig. 1).
back to the atmosphere, as the water flows downstream Due to the complexity of the hydraulic jump phe-
(Chanson, 2011). Another determinant characteris- nomenon, many of its characteristics are still not
tic of hydraulic jumps is the presence of turbulence, completely understood, which explain the need of con-
causing intense fluctuations of flow parameters like tinuous studies in this theme. Some of the issues that
velocity, surface position and pressure. Regarding the have been more recently addressed are: 1. numeri-
longitudinal position of the hydraulic jump, the front cal simulation of the flow (Chern & Syamsuri 2013,
of the jump moves back and forward in a very chaotic Simões et al. 2010) 2. influence of the bed rough-
movement. Similarly, the flow in the vertical direc- ness on flow patterns (Carollo et al. 2007, Ead &
tion changes rapidly in time, producing many splashes Rajaratnam 2002) 3. performance of baffle blocks
and droplets, which reach heights that may exceed the (Habibzadeh et al. 2012, 2014) 4. air-water interface
value of the vertical dimension of the jump itself. characteristics (Murzyn et al. 2007, Nóbrega et al.
The water surface mean profile (evolution of the 2014, Wang et al. 2015) 5. properties of the air-
mean water depth with the longitudinal distance) water flow (two-phase flow) – (Chanson 2011, Zhang
may be viewed as being composed firstly by a “sat- et al. 2014) 6. determination of the turbulent field
uration exponential” profile which then rises until using PIV (particle image velocimetry), BIV (bub-
the condition of uniform flow (Lj – hydraulic jump ble image velocimetry) and ADV (acoustic Doppler
length). In the first region, the flow usually presents a velocimeter) – (Lennon & Hill 2006, Lin et al. 2012,

629
Liu et al. 2004, Mignot & Cienfuegos 2011) 7. the- probes and the acoustic sensors, integral turbulent time
oretical and semi-empirical formulations (Beirami & and length scales, and Strouhal numbers. The study
Chamani 2010, Schulz et al. 2015, Valiani 1997). of Nóbrega et al. (2014) showed a good comparison
Regarding the pressure field generated by hydraulic between mean free surface level from ultrasonic sen-
jumps, an expressive number of researchers carried out sor measurements and images, which were obtained
experimental investigations, some of them devoted to using a high speed camera focusing the flow from the
describe the evolution of the pressure along the bot- sidewall.
tom of the channel. The first papers on this theme were Because of the influence of the strong turbulence on
probably published by Vasiliev & Bukreyev (1967), both the deformation and breakup of the free surface
Wisner (1967), and Abdul Khader & Elango (1974). and on the pressure fluctuations in hydraulic jumps,
Abul & Elango (1974) used pressure cells on the bot- it is expected that the statistical quantities of the free
tom of the channel for three series of experiments in surface fluctuations in hydraulic jumps may be cor-
order to measure mean pressures and their fluctuations related to those of the pressure fluctuations on the
in the center line of the jump. bottom of the flume. However, there is not a study
Bowers & Toso (1988) also presented information that provides this comparison and quantitative analy-
about pressure fields in order to analyze the possi- ses for this relationship. Therefore, this paper aims to
ble causes of failure of the Karnafulli spillway. Based furnish statistical information about the free surface
on their experiments in a physical model, they dis- position and the pressure field along hydraulic jumps,
cussed that the slab failure was probably caused by the comparing the free surface measurements using ultra-
differences in fluctuating pressure at the chute of the sonic sensor and the pressure data from the literature.
spillway and the chute slab. Additional information of The results may conduce to indirect evaluations of
maximum and minimum instantaneous pressure distri- the behavior of mean statistical characteristics of the
bution along the flow, in relation to mean pressure, was pressure evolution based on observations of the free
further presented by Toso & Bowers (1988). Moreover, surface.
they evaluated the influence of different physical and
flow conditions (inflow Froude number, developed and
undeveloped incident flow, chute slopes, chute blocks, 2 EXPERIMENTAL SETUP
intermediate blocks and end sill) on the pressure field.
Fiorotto & Rinaldo (1992) also developed studies 2.1 Facilities and flow conditions
on fluctuating pressure in hydraulic jumps in view of The experiments about the evolution of the surface
the stability of the stilling basin. The paper gives infor- were conducted in two recirculating channels: one at
mation about statistical parameters along hydraulic the Hydraulics Laboratory of the Engineering School
jumps, covering a broader range of inflow Froude num- of São Carlos – University of São Paulo, Brazil;
bers than the previous works. Armenio et al. (2000) and the other at the Undergraduate Laboratory of
also evaluated statistical quantities of pressure fluctu- the University of Alberta, Canada. The channels are
ations at the bottom of a hydraulic jump over a negative identified here as 1 and 2, respectively, as shown in
step, furnishing results of extreme values and spa- Figure 2.
tial correlation structures. Studies on different aspects Channel 1 is 41 cm wide, 60 cm height and has bed
of pressure fluctuations and their decaying, pressure and sidewalls built in concrete. The flow rate was mea-
distribution, among other relevant information about sured by a triangular weir, displaced downstream of
the pressure fields were conducted by Marques and the water tank. Hydraulic jumps were produced down-
coworkers, like: Marques et al. (1997), Marques et al. stream of a broad crested weir of 24 cm height and
(2000), Neto & Marques (2008). 46 cm long.
In relation to the position of the free surface in Channel 2 is 50 cm wide, 5 m long and has plex-
hydraulic jumps, some authors have focused their iglass side walls and an aluminum bed. The flow
attention on the investigation of the dynamics of the rate was measured by a magnetic flow transmitter
air-water interface, viewed as dependent of interac- (Foxboro® Model IMT25) connected to the recircu-
tions between the large-scale eddies and the free lation pipe. The supercritical condition was formed
surface. Murzyn et al. (2007) worked on the theme as the water passed under a sluice gate at the flume
and used wire gauges in their experimental studies. entrance, with an opening of 2.8 mm.
Besides determining the free surface and relevant tur- The position of the hydraulic jump in both channels
bulence profiles, they furnished free surface length was controlled by a tail gate. The mean front of the
scales for both longitudinal and the transversal direc- hydraulic jump was positioned at the maximum water
tions. Kucukali & Chanson (2008), Chachereau & jet contraction downstream of the broad crested weir or
Chanson (2011), and Murzyn & Chanson (2009) per- sluice gate. Seven hydraulic jumps with inflow Froude
formed measurements using acoustic displacement numbers from 2.8 to 5.3 were tested. Table 1 presents
meters, described as a non-intrusive technique. Simul- the flow conditions.
taneously, two-phase flow properties were recorded In all studies the possibility of scaling is always
using phase detection probes. Their results include: a present question. According to Murzyn & Chanson
free surface profiles, free surface fluctuations, spectral (2008), for hydraulic jumps with Re1 up to 1 × 105 ,
analysis of the data obtained with the phase detection the rate of entrained air and air-water interfacial area

630
Figure 2. Experimental apparatus: a) Flume 1: broad crested
weir; b) Flume 2: sluice gate.

Table 1. Experimental flow conditions.

Flume 1
Broad crested Weir
Figure 3. Ultrasonic sensor in a displacement track (detail).
Q y1 y2 V1
Code L/s cm cm m/s Re1 F1
by Simões et al. (2010, 2013), in their studies on
II-21 21.0 2.5 14.2 2.04 51,000 4.11 hydraulic jumps and stepped spillways.
II-31 31.0 3.5 17.4 2.15 75,250 3.67 The ultrasonic sensor was mounted in a displace-
II-40 40.0 5.0 21.4 1.95 97,500 2.78
ment track at the central longitudinal axis of the flume,
Flume 2
above the free surface, in order to determine instanta-
Sluice gate neous positions of the surface (Fig. 3). The sampling
time and frequency for the measurements at each posi-
Q y1 y2 V1 tion were 2.0 minutes and 25.0 Hz, respectively. The
Code L/s cm cm m/s Re1 F1 range of distance measured by the sensor is from
15 to 600 cm, with a resolution of 1 mm. Accord-
III-21 20.9 2.8 10.6 1.56 43,680 2.98 ing to the manufacturer, the emitted sound waves
III-27 26.9 2.9 14.2 1.94 56,260 3.64 travel at a speed of about 343 m/s, forming a coni-
III-34 34.2 2.9 17.9 2.44 70,760 4.58 cal frustum with angle between 15 and 20 degrees and
III-39 38.9 2.9 20.5 2.80 81,200 5.26 smaller base diameter of 3.7 cm, corresponding to the
senser/detector. The sampling was accomplished by a

Q = flow rate; y1 = supercritical depth; y2 = subcritical
depth; V1 = supercritical velocity; Re1 = supercritical
Logger Lite – Vernier® software. The ultrasonic sensor
Reynolds number; F1 = supercritical Froude number. records the time for the sound wave emitted by the sen-
sor (in a conical frustum region) to return to the device,
after reaching the air-water interface or its splashes.
This travel time is proportional to the distance between
are underestimated. The authors also discuss that the sensor and the obstacle, hence instantaneous posi-
the dynamic similarity of two-phase flow cannot be tions of the obstacles are registered, being the water
achieved unless working in a full scale 1:1. Teixeira depths statistically determined in the sequence. Details
et al. (2012), and Teixeira (2008) conducted pressure about the statistical procedures followed to obtain the
measurements in spillways, for the Porto Colômbia mean water depths at each measurement position may
Dam in the prototype and in scales 1:32, 1:50, 1:80, be found in Simões (2012) and Nóbrega (2014).
1:100. Their investigation showed that the mean pres-
sure behaviors could be studied in the 1:100 scale
without loss of information. Because pressure and 3 RESULTS
depths are correlated, but also air entrainment may
affect this correlation, future investigations regarding 3.1 Free surface profile
the scale effects are still welcomed.
Although the free surface profile of hydraulic jumps
are very unstable, both in horizontal and vertical direc-
tions, due to the jump toe oscillations and high level
of turbulence, a mean profile can be defined from the
2.2 Instrumentation
time series for each location obtained by the ultrasonic
An ultrasonic sensor (Vernier® – Go!Motion) was sensor. The time series in the experiments were com-
used to detect the free surface position along the posed by 3000 data for each position. Figure 4 shows
hydraulic jumps. This sensor was also previously used the mean depths along a hydraulic jump for F1 = 3.64,

631
Figure 5. Mean pressure distribution along hydraulic jump.
Marques et al. (1997).
Figure 4. Instantaneous depths and mean profiles for
hydraulic jump with F1 = 3.64.

the points representing the instantaneous values, and


blue curves representing the mean value ±3 times
the standard deviation. As can be observed in Figure
4, the standard deviation increases within the roller
length (x − x1 ≈ 50, defined by visual observations),
and gradually decreases afterwards. The maximum
depth (y = 16.5 at x − x1 ≈ 60), considering the mean
value plus 3 times the standard deviation, is about
16% higher than the subcritical depth y2 = 14.2, for Figure 6. Mean, maximum and minimum pressures along
F1 = 3.64. hydraulic jump. Marques et al. (1997).
In order to define a representative curve for all
experimental data, the free surface profiles were plot-
ted in a nondimensonal graph. Figure 7a shows that
for hydraulic jumps with F1 = 4.58 and 5.26, the free
surface profile presented a wavy shape. A possible Additionally, multiple obstacles (drops) may generate
cause may be the proximity of the jump toe to the reflections between them before attaining the detec-
sluice gate. For the other inflow Froude numbers tested tor, implying in larger travel times, and, consequently,
here, the mean surface profiles of the hydraulic jumps larger calculated distances. Drops are caused by the
are smooth and gradually rise until the uniform flow strong oscillations of the jump in the longitudinal and
condition (and further decrease accordingly the flow vertical directions, so that the sound in the conical
conditions). frustum may be subjected to the mentioned reflec-
Some authors have proposed experimental solu- tions. The maximum extreme values form a cloud of
tion for the free surface. Hager (1993), for example, points which is more dense until about 1.5 times the
defined an expression for the dimensionless depth as dimensionless subcritical depth (y − y2 )/(y2 − y1 ) = 1,
a function of the dimensionless length x/Lr . Other occurring in the range of positions x/(y2 − y1 ) ≈ 2 to
expressions for free surface profiles were proposed by 6, approximately. The maximum differences between
Schulz et al. (2015), which were adjusted to the same maximum extreme depths and mean depths were found
experimental data presented here. at x/(y2 − y1 ) ≈ 1. Comparing these behaviors and val-
Visually comparing the mean depths to the mean ues with the results of Figure 6, which shows the max-
pressures along hydraulic jumps (Figures 7a and 5), imum, minimum and mean dimensionless pressure,
similar main patterns are observed. However, the curve we note that downstream x/(y2 − y1 ) ≈ 8, the extreme
of mean pressure attains (Px − y2 )/(y2 − y1 ) ≈ 1 near values of depth and pressure become approximately
the position x/(y2 − y1 ) ≈ 8, while the curve of mean constant. The maximum pressure events concentrate
depth attains the value of (y − y2 )/(y2 − y1 ) ≈ 1 at around x/(y2 − y1 ) ≈ 2. The behaviors of the events
x/(y2 − y1 ) ≈ 6 without considering the two oscillat- of minimum pressure and minimum depths, although
ing surfaces. If considering the oscillating cases, the of course located under the points representing the
value x/(y2 − y1 ) ≈ 8 is also observed. mean values of both variables, evolve differently con-
Figure 7b presents the mean and extreme depth val- cerning the length of the concavity of the subjacent
ues for the tested flow conditions. It can be observed curve. One reason of this difference was already men-
that near the jump toe, some negative values of water tioned, that is, depths have no negative values (negative
depths are recorded, which are unreal. Negative values measurements may be related to the conical frustum
may be caused by different positions in the measure- measurement region and to reflections of the sound in
ment cone region. Considering an angle of 20 degrees it). Another reason may be the fact that the pressure
for the cone, distances up to 1.5% larger than the ver- field is affected by the flow separation at the flume
tical height may be registered for horizontal surfaces. bottom.

632
Figure 7. a) Mean free surface profile; b) Minimum and maximum depths recorded using ultrasonic sensor; c) Fluctuation
coefficient depth distribution; d) Skewness depth distribution along the hydraulic jump.

633
3.2 Fluctuation coefficient
The fluctuation coefficient relates the mean standard
deviation of the depth data set at a certain position – Cy′
(or pressure data set – Cp′ ) to the inflow kinetic energy
at the toe of the jump.

Figure 8. Fluctuation coefficient pressure distribution


where σy = standard deviation of the depth data set; along hydraulic jump. Adapted from Neto & Marques (2008),
and Lopardo (2012).
σp = standard deviation of the pressure data set;
V1 = inflow velocity; and g = gravity.
Towards the tail of the jump, the fluctuation coeffi-
cient presented approximately a constant value, around
0.02 (Fig. 7c). In general, the maximum value of the
fluctuation coefficient is about 0.08, at the position
(x − x1 )/(y2 − y1 ) ≈ 1. One exception was however
observed: experiment I 40, with F1 = 2.78, presented
a different behavior, with the maximum value of 0.13.
These results are similar to those presented by Neto
& Marques (2008), and Lopardo (2012), for pressure
fluctuations (Fig. 8). The maximum fluctuation coeffi-
cient of Neto & Marques (2008) experiments, based on
the mentioned pressure measurements (using pressure Figure 9. Skewness pressure distribution along hydraulic
transducers), were of about 0.08 at position x/y1 ≈ 10, jump. Adapted from Neto & Marques (2008), and Lopardo
for hydraulic jumps with F1 = 4.55 and 5.01. Accord- (2012).
ing to Lopardo (2012) measurements, the maximum
values of Cp′ were also around 0.06–0.07, at position
(x − x1 )/(y2 − y1 ) ≈ 2.
For both variables (depth and pressure), down- before the position (x − x1 )/(y2 − y1 ) ≈ 5, i.e., the val-
stream the position (x − x1 )/(y2 − y1 ) ≈ 10, the coef- ues tend to be higher than the mean value. This may
ficients become approximately constant. The value be also extracted from Figure 4, in which instanta-
obtained for depth measurements is about 0.02 (see neous values at the roller region tend to be higher than
Fig. 7c), which however is higher than the value the curve representing the mean depth + 3 times the
of 0.01 found by Neto & Marques (2008), Lopardo standard deviation.
(2012), and Abdul Khader & Elango (1974), for the The maximum skewness values were observed
decay of the standard deviation of the pressure along near the jump toe. In the downstream direction,
the jump (Fig. 8). This is an interesting conclusion, the skewness values stabilize around zero. This
because it evidences that both measurements (depth trend is similar to data of Neto & Marques (2008),
and pressure) may be used to limit the length of the and Lopardo (2012) for pressure measurements,
hydraulic jump. Because asymptotic behaviors are as shown in Figure 9. However, downstream the
observed, conditional definitions of lengths may be position (x − x1 )/(y2 − y1 ) ≈ 4, the skewness values
used involving mean values or extreme values. attain a minimum negative value around the posi-
tion (x − x1 )/(y2 − y1 ) ≈ 6, tending then to zero down-
stream of (x − x1 )/(y2 − y1 ) ≈ 8. As discussed by
3.3 Skewness Lopardo & Casado (2007), the transition from positive
to negative skewness values seems to imply the possi-
The skewness is the third central moment and is a quan- bility of a boundary layer separation from the stilling
tity that represents the asymmetry of the probability basin bottom.
distribution of the measured instantaneous values in The present results show that there are remark-
relation to the mean value. able similarities between the general behavior of the
statistical variables associated to depth and pressure
measurements in hydraulic jumps. When consider-
ing details as the length of concavity regions and
the presence of critical points (minima, for example),
where yi = instantaneous depth at location x; yx = mean differences were also observed. The similarities are
value at location x; n = length of the sample data; related to the same cause of both fluctuations (pres-
σx = standard deviation at location x. sure and depth), that is, the turbulent movement of
As can be observed from Figure 7d, for the depth the flow in the hydraulic jump. The differences follow
measurements, the skewness distribution is positive from the different nature of the physical parameters

634
being analyzed, which point that more studies are nec- F1 Inflow Froude number
essary to verify possible quantitative relations between P Pressure at location x
them. A goal for such a relationship would be to use Q Flow rate
one of the measurements as indirect “quantification” Re1 Inflow Reynolds number
of the second. Results of laboratory flumes were used V1 Velocity of the flow at the supercritical
in the present study, involving small dimensions and position
flow rates. The observed similarities suggest to extend X Longitudinal distance along hydraulic jump
the study to larger dimensions, so that eventual scale x1 Jump toe position
effects may also be also evidenced. Y Water depth at location x
y1 Supercritical depth
y2 Subcritical depth
σy , σp Standard deviation of water depth data and
4 CONCLUSIONS
pressure data at location x, respectively
Ultrasonic sensors have been employed in hydraulic
jumps for measurement of flow depths. It has proved
to be a simple and promising technique for studies of ACKNOWLEDMENTS
strongly turbulent open flows. Because the flow depth
fluctuations and pressure fluctuations are caused by The authors thank CAPES, CNPq and Fapesp for fund-
the same turbulent condition, it is expected they are ing this study. The authors acknowledge Dr. David
also related to each other. In the present study, results Z. Zhu for supervising the student Juliana during her
of experiments conducted for different inflow Froude stay at the University of Alberta, and the technician
numbers and upstream conditions (broad crested weir Perry Fedun, for helping during the experiments in the
and sluice gate), using ultrasonic sensors, were pre- University of Alberta.
sented. The sensors were moved along the central line
of the experimental flumes, and were used to provide
statistical information about the surface evolution. Sta- REFERENCES
tistical parameters of the flow depths (mean values,
standard deviation, skewness) were compared to sta- Abdul Khader, M.H. & Elango, K. 1974. Turbulent pres-
sure field beneath a hydraulic jump. Journal of Hydraulic
tistical values of pressure fields obtained from the
Research 12(4): 469–489.
literature. Armenio, V., Toscano, P. & Fiorotto, V. 2000. On the effects
The statistical treatment of the pressure fluctuations of a negative step in pressure fluctuations at the bottom of
in the literature showed results having behaviors simi- a hydraulic jump. Journal of Hydraulic Research 38(5):
lar to that of the flow depths. The values of the standard 359–368.
deviations of both fluctuations (depth and pressure) Beirami, M.K. & Chamani, M.R. 2010. Hydraulic jump in
reached their maxima in the region of the jump roller, sloping channels: roller length and energy loss. Canadian
decaying then to a reasonable constant value after- Journal of Civil Engineering 37: 535–543.
wards, when the flow turbulence is less intense, and Bowers, C.E. & Toso, J.W. 1988. Karnafuli project, model
studies of spillway damage. Journal of Hydraulic Engi-
the flow stablishes the subcritical condition.
neering 114(5): 469–483.
The present results involve relatively small dimen- Carollo, F.G., Ferro, V. & Pampalone, V. 2007. Hydraulic
sions and flow rates, but the mentioned similarities jumps of rough beds. Journal of Hydraulic Engineering
point to the convenience of more studies for larger 133(9): 989–999.
dimensions, evidencing scale effects, considering that Chachereau,Y. & Chanson, H. 2011. Free surface fluctuations
phenomena of different natures are involved in the data and turbulence in hydraulic jumps. Experimental Thermal
for depths and for pressure. and Fluid Science 35(6): 896–909.
This study also allows to suggest that the length of Chanson, H. 2011. Bubbly two-phase flow in hydraulic
hydraulic jumps (and eventually also of the rollers) jumps at large Froude numbers. Journal of Hydraulic
Engineering 137(4): 451–460.
may be quantified based on the decay of the fluc-
Chern, M. & Syamsuri, S. 2013. Effect of corrugated bed on
tuations of the flow depths or the pressures. These hydraulic jump characteristic using SPH method. Journal
characteristic lengths are of difficult measurement by of Hydraulic Engineering 139: 221–232.
visual observation, pointing to the necessity of quanti- Ead, S.A. & Rajaratnam, N. 2002. Hydraulic jumps on cor-
tative criteria for them, which depend on the equipment rugated beds. Journal of Hydraulic Engineering 128(7):
used to obtain the experimental data. 656–663.
Fiorotto, V. & Rinaldo, A. 1992. Turbulent pressure fluc-
tuations under hydraulic jumps. Journal of Hydraulic
SYMBOLS Research 30(4): 499–520.
Habibzadeh, A., Loewen, M.R. & Rajaratnam, N. 2012. Per-
Cy′ , Cp′ Fluctuation coefficient of water depth and formance of baffle blocks in submerged hydraulic jumps.
Journal of Hydraulic Engineering 138(10): 902–908.
pressure data at location x, respectively Habibzadeh, A., Loewen, M.R. & Rajaratnam, N. 2014.
Lr Roller length Mean flow in a submerged hydraulic jump with
Lj Hydraulic jump length baffle blocks. Journal of Engineering Mechanics,
Ltr Hydraulic jump transition length 10.1061/(ASCE)EM.1943-7889.0000713, 04014020.

635
Hager, W.H. 1993. Classical hydraulic jump: free surface Schulz, H.E., Nóbrega, J.D., Simões, A.L.A., Schulz, H.
profile. Canadian Journal of Civil Engineering 20(3): & Porto, R.M. 2015. Details of Hydraulic Jumps for
536–539. Design Criteria of Hydraulic Structures, Hydrodynamics-
Kucukali, S. & Chanson, H. 2008. Turbulence measure- Concepts and Experiments, Schulz, H.E. (ed.), InTech,
ments in the bubbly flow region of hydraulic jumps. Available from: http://www.intechopen.com/books/hydro
Experimental Thermal and Fluid Science 33: 41–53. dynamics-concepts-and-experiments/details-of-hydraulic-
Lennon, J.M. & Hill, D.F. 2006. Particle image velocimetry jumps-for-design-criteria-of-hydraulic-structures.
measurements of undular and hydraulic jumps. Journal of Simões, A.L.A. 2012. Escoamentos turbulentos em canais
Hydraulic Engineering 132(12): 1283–1294. com o fundo em degraus: resultados experimentais,
Lin, C., Hsieh, S-C., Lin, I-J., Chang, K-A. & Raikar, R. 2012. soluções numéricas e proposições teóricas, PhD Thesis,
Laboratory measurements of a steady breaker using PIV School of Engineering São Carlos, University of São
and BIV. Coastal Engineering Proceedings 33: 1–9. Paulo, Brazil. Available from: http://www.teses.usp.br/
Liu, M., Rajaratnam, N. & Zhu, D.Z. 2004. Turbulence struc- teses/disponiveis/18/18138/tde-01092014-090055/en.php.
ture of hydraulic jumps of low Froude numbers. Journal [in Portuguese]
of Hydraulic Engineering 130(6): 511–520. Simões, A.L.A., Schulz, H.E. & Porto, R.M. 2010. Sim-
Lopardo, R.A. 2012. Internal flow of free hydraulic jump ulação numérica e verificação experimental da posição
in stilling basins. In 4th IAHR International Sympo- da superfície livre de um ressalto hidráulico em um
sium on Hydraulic Structures, 9–11 February 2012. Porto canal retangular. In XXIV Congreso Latinoamericano de
Portugal: IAHR. Hidráulica, November 2010. Punta Del Este, Uruguay:
Lopardo, R.A. & Casado, J.M. 2007. Boundary layer sep- IAHR. [in Portuguese]
aration beneath submerged jump flows. In XXXII IAHR Simões, A.L.A, Schulz, H.E., Porto, R.M. & Gulliver, J.S.
Congress. Venice, Italy. 2013. Free-surface profiles and turbulence characteristics
Marques, M.G., Drapeau, J. & Verrette, J-L. 1997. Flutuação in skimming flows along stepped chutes. Journal of Water
de pressão em um ressalto hidráulico. Revista Brasileira Resource and Hydraulic Engineering 2(1): 1–12.
de Recursos Hídricos 2(2): 45–52. [in Portuguese] Teixeira, E.D. 2008. Efeito de escala na previsão dos
Marques, M.G., Gomes, J.F. & Endres, L.A.M. 2000. valores extremos de pressão junto ao fundo em
Oscilação da posição inicial do ressalto hidráulico e o bacias de dissipação por ressalto hidráulico, PhD The-
campo de pressões na soleira de bacias de dissipação. sis, Institute of Hydraulics Research, Federal Univer-
In Simpósio de Recursos Hídricos do Nordeste, 21–24 sity of Rio Grande do Sul, Brazil. Available from:
November 2000. Natal, Brazil: ABRH [in Portuguese] http://hdl.handle.net/10183/17516. [in Portuguese].
Mignot, E. & Cienfuegos, R. 2011. Spatial evolution of tur- Teixeira, E.D., Dai Prá, M., Wiest, R.A. & Marques, M.G.
bulence characteristics in weak hydraulic jumps. Journal 2012. Efeitos de escala nos valores de pressão média
of Hydraulic Research 49(2): 222–230. junto ao fundo em bacias de dissipação por ressalto
Murzyn, F. & Chanson, H. 2008. Experimental assessment hidráulico submergido. Revista Brasileira de Recursos
of scale effects affecting two-phase flow properties in Hídricos 17(2): 87–100. [in Portuguese]
hydraulic jumps. Experiments in Fluids 45(3): 513–521. Toso, J.W. & Bowers, C.E. 1988. Extreme pressures in
Murzyn, F. & Chanson, H. 2009. Free-surface fluctuations in hydraulic jumps stilling basins. Journal of Hydraulic
hydraulic jumps: Experimental observations. Experimen- Engineering 114(8): 829–843.
tal Thermal and Fluid Science 33: 1055–1064. Valiani, A. 1997. Linear and angular momentum conservation
Murzyn, F., Mouzé, D. & Chaplin, J.R. 2007. Air-water inter- in hydraulic jump. Journal of Hydraulic Research 35(3):
face dynamic and free surface features in hydraulic jumps. 323–354.
Journal of Hydraulic Research 45(5): 679–685. Vasiliev, O.F. & Bukreyev, V.I. 1967. Statistical characteristics
Neto, E.F.T.N. & Marques, M.G. 2008. Análise do campo de of pressure fluctuations in the region of hydraulic jump.
pressões em ressalto hidráulico submergido a jusante de In 12th Congress of the International Water Association
uma comporta. Revista Brasileira de Recursos Hídricos for Hydro-Environment Engineering and Research, 11–14
13(4): 55–68. [in Portuguese] September 1967. Colorado, USA: IAHR.
Nóbrega, J.D. 2014. Metodologia teórica e experimental para Wang, H., Murzyn, F. & Chanson, H. 2015. Interaction
determinação das características do ressalto hidráulico between free surface, two-phase flow and total pressure in
clássico, MScDissertation, School of Engineering at São hydraulic jumps. ExperimentalThermal and Fluid Science
Carlos, University of São Paulo, Brazil, Available from: 64: 30–41.
http://www.teses.usp.br/teses/disponiveis/18/18138/tde- Wisner, P. 1967. On the bottom pressure pulsations of the
30092014-105847/pt-br.php. [in Portuguese] closed conduit and open channel hydraulic jumps. In
Nóbrega, J.D., Schulz, H.E. & Zhu, D.Z. 2014. Free surface 12th Congress of the International Water Association for
detection in hydraulic jumps through image analysis and Hydro-Environment Engineering and Research, 11–14
ultrasonic sensor measurements. In Hubert Chanson and September 1967. Colorado, USA: IAHR. [in French]
Luke Toombes (ed.), Hydraulic structures and society – Zhang, W., Liu, M., Zhu, D.Z. & Rajaratnam, N. 2014. Mean
Engineering challenges and extremes. 5th IAHR Interna- and turbulent bubble velocities in free hydraulic jumps
tional Symposium on Hydraulic Structures, 25–27 June for small to intermediate Froude numbers. Journal of
2014. Brisbane Australia, doi:10.14264/uql.2014.42. Hydraulic Engineering 140(11): 04014055 1–9.

636

You might also like