(Asce) Hy 1943-7900 0001669

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 3

Discussions and Closures

Closure to “Energy Dissipation of a Type III Basin under


Design and Adverse Conditions for Stepped
and Smooth Spillways” by D. Valero,
D. B. Bung, and B. M. Crookston
D. Valero, Ph.D. subcritical flows the pressure term will be larger than the velocity
Lecturer, River Basin Development Chair Group, Water Science and En- term, whereas the velocity term is larger than the pressure term for
gineering Dept., IHE Delft Institute for Water Education, Delft, 2611 AX, supercritical flows. Thus, keeping constant the left side of Eq. (1),
Downloaded from ascelibrary.org by 105.106.9.188 on 12/12/19. Copyright ASCE. For personal use only; all rights reserved.

Netherlands. Email: d.valero@un-ihe.org while decreasing D2 implies that larger Fx (stilling basin response)
is required to keep the jump located at a given position. Hence, a
D. B. Bung, Ph.D. simplified momentum-based analysis suggests that lowering the
downstream tail water level directly represents adverse flow
Professor, Hydraulic Engineering Section, Civil Engineering Dept., FH
Aachen Univ. of Applied Sciences, Aachen 52066, Germany. Email: conditions for the stilling basin, thus endorsing the validity of the
bung@fh-aachen.de approach of the original paper.
The discussers added some further comments on the data extrac-
B. M. Crookston, Ph.D., P.E., M.ASCE tion from the numerical model as well as the general model setup.
The writers also provide some contextual information regarding the
Assistant Professor, Dept. Civil and Environmental Engineering, Utah request to report pressure distributions from Reynolds-averaged
Water Research Laboratory, Utah State Univ., 8200 Old Main Hill, Logan
Navier-Stokes (RANS) eddy viscosity models to assist practi-
UT 84322-8200 (corresponding author). ORCID: https://orcid.org/0000
-0003-1259-8540. Email: brian.crookston@usu.edu tioners on the design of stilling basins against high-pressure loads.
To begin, the mean pressure at a given point (gravity effect included
within the term p̄) can be estimated as (Pope 2000, p. 85)
https://doi.org/10.1061/(ASCE)HY.1943-7900.0001482
1 ∂ ū ∂ ūj ∂ 2 ui0 uj0
The writers appreciate the discussers’ interest in this paper and − ∇2 p̄ ¼ ρ i þ ð2Þ
agree that further research is needed if standardized design guide- ρ ∂xj ∂xi ∂xi ∂xj
lines are to be developed to account for the effects of a stepped
chute invert on hydraulic jump stilling basin performance. This where ūi and ui0 = mean and fluctuating parts of the velocity field
was indeed the ultimate motivation of the original paper. The writ- and xi the cartesian coordinates. Subscripts denote the correspond-
ers are grateful for the opportunity to respond to several comments ing coordinates (1, 2, 3 for x, y, z). The right side of Eq. (2) includes
noted by the discussers while providing additional thoughts and two main sources that can result in deviations from hydrostatic
perspective on computational fluid dynamics (CFD) modeling pressures, namely changes in flow trajectories, and the distribution
and hydraulic jump stilling basin design. of Reynolds stresses. While changes in flow trajectories are one of
The discussers noted the linkage between the lowering tailwater the weakest aspects of the RANS models (see roller length predic-
levels and the adverse flow conditions for the stilling basin perfor- tions accuracy), Reynolds stresses are not directly computed or not
mance. Indeed, a hydraulic jump stilling basin has a design flow even incorporated (in the case of eddy viscosity models). Note that
and suggested minimum tailwater, yet performance should be Eq. (2) allows the computation of the mean pressure, yet the pres-
checked over the anticipated operation range including a sensitivity sure field is clearly transient, and the computation of peak pres-
analysis of backwater effects. Some uncertainty may exist in sures, of greater interest for practitioners, would require further
backwater estimations provided by the downstream channel that, considerations. Specifically, RANS equations provide approximate
depending on the numerical river model and input information, time-averaged solutions (yet account for some unsteadiness) and it
may exceed 1 m for various flood events. Thus, a practitioner would be prudent to compare the computed pressure field time
is interested in the location and stability of the hydraulic jump series from a RANS model to reliable experimental or field data.
for a range of tailwater elevations. Streamwise momentum balance Forasmuch as the simplest case of the classic hydraulic jump has
can thereby shed some light on adverse flow conditions. By assum- served as a RANS modeling workbench, validation has to our
ing uniform velocity distributions and taking cross sections where knowledge not been conducted for neither mean nor turbulent pres-
hydrostatic flow conditions are re-established, this balance can be sures [see recent review of Viti et al. (2019)]. This step may be
written as addressed first. Moreover, little is known on how the aeration
will impact both terms of the right side of Eq. (2), and thus a
1 1 conservative analysis was preferred so as to not mislead the reader
ρgBD21 þ ρQV 1 ¼ ρgBD22 þ ρQV 2 þ Fx ð1Þ
2 2 with the results. The original paper compared numerical results to
the important study of Frizell and Svoboda (2012), who did not
where subscripts 1 and 2 refer to upstream and downstream sec- provide pressure field experimental data. Therefore, we are unable
tions, respectively; ρ = water density; g = gravity acceleration; to perform a validation assessment of the predicted average pres-
B = channel width; D = flow depth; V = mean flow velocity; sures because quantitative information regarding numerical results
Q = volumetric flow rate; and Fx is a drag force opposing the flow. and model accuracy cannot be provided without this experimental
For short, the term related to D will be named the pressure term data. Caution must prevail when intensive validation has not been
and the term related to V will be the velocity term. Note that in reported by the community and little physical evidences hold.

© ASCE 07019015-1 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(2): 07019015


Downloaded from ascelibrary.org by 105.106.9.188 on 12/12/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Folsom Dam auxiliary spillway looking upstream at the stepped chute with Type III hydraulic jump stilling basin and supercavitating blocks.
(Image courtesy of US Army Corps of Engineers.)

However, some information is available in the literature for hy- design, e.g., for sidewall heights (Hunt and Kadavy 2017).
draulic jump stilling basins of a different geometry. Meireles et al. Notwithstanding its conspicuous occurrence in prototypes and
(2010) reported pressure head results in a Type III stilling basin. many scaled physical models, air entrainment (and detrainment for
Some additional information on pressures for stilling basin flows that matter) is oftentimes neglected in analytical analysis, yet ac-
can be found in Bung et al. (2012). The effect of air entrainment on curate flow descriptions have been obtained (e.g., Castro-Orgaz and
maximum pressures in plunge pools is discussed by Bollaert and Hager 2017, pp. 481–485). A guiding principal for numerical
Schleiss (2003), illustrating a challenge with model scale with re- analyses is, for a given objective, determining the appropriate
gards to two-phase flow pressure fields. Hunt and Kadavy (2018) model and corresponding complexity. The original paper refer-
provide preliminary results of a large-scale stilling basin study enced numerous instances in the literature where investigators
equipped with pressure transducers limited to the basin centerline. successfully employed a single-phase flow model, and the paper
Additionally, the discussers state that the employed single-fluid provides insights into the performance of this approach for a Type
approach may have some effects on energy dissipation results and III basin for consideration. A single-fluid approach should be bal-
stilling basin performance as the entrained air phase is ignored. To anced against prediction accuracy and computation cost. Specific
overcome this issue, a two-fluid modeling approach is recom- considerations must be given to the computational mesh and resolv-
mended. We agree with the discussers on the aerated nature of ing scales related to air entrainment and sufficient resolution of air
the flow, but nonetheless the RANS modeling community has bubbles while avoiding large numerical diffusion, etc. Approaches
achieved relevant outcomes for the classic hydraulic jump case both to include air effects merit greater insight on the flow behavior to
with single-phase and multiphase modeling. Moreover, entrain- formulate accurate models for highly aerated turbulent flows, such
ment of big voids can still be captured by the model even if not as those occurring in spillways and stilling basins. In many instan-
at all sizes. Please note that given the air–water density difference, ces, this additional complexity represents an unjustified increase in
the impact on the overall momentum balance [Eq. (1)] may remain computer power while introducing additional uncertainty into the
low. When considering selection of either a single-phase or multi- model results by considering air entrainment, which in this case
phase numerical model, a relevant question is whether for the de- might be larger than the deficiencies observed in analytical analyses
sired outcome, it is worth solving two sets of equations; i.e., will where it was not considered at all. Further knowledge on air–water
solving these equations with the necessary model details including flows behavior may assist scientists on formulating robust multi-
a sufficiently small scale be more accurate than not solving them? phase models, but at this point we believe that numerical modeling
And, if the air effect is not completely reproduced, should we of air entrainment is not at a mature stage.
use subgrid models to approximate its effect on the main flow? Note The writers hope that the community will continue to investigate
that the latter includes use of numerous semiempirical equations complex flows in spillways, including various types of hydraulic
and flow conceptualizations for which little knowledge is available jump stilling basins downstream of stepped chutes (Fig. 1). Addi-
on their validity out of the calibration range [e.g., calibration- tional insights from numerical modeling, laboratory studies, and
validation of Valero and García-Bartual (2016) and issues field observations are needed. This would provide further under-
reported by Valero and Bung (2015) for a different spillway]. standing into their complex flow behaviors and hydraulic structure
Addressing these questions would be an interesting contribution design guidance for practitioners, who balance a variety of project
beyond the scope of the original paper. Flow bulking and also requirements against hydraulic performance including project
splash and spray from the turbulent flows conveyed through a economy, constructability, and site-specific conditions ranging
spillway should be considered for spillway and stilling basin from site geology, basin hydrology, ecological impacts, and project

© ASCE 07019015-2 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(2): 07019015


risks to public safety, emergency response, and even unauthorized Hunt, S. L., and K. C. Kadavy. 2018. “Preliminary results for embankment
recreational activities occurring in spillways. dam stepped spillway stilling basin research.” In Proc., 7th IAHR Int.
Symp. on Hydraulic Structures, edited by D. Bung and B. Tullis.
Logan, UT: Utah State Univ.
References Meireles I., Matos J., and A. Silva-Afonso. 2010. “Flow characteristics
along USBR type III stilling basins downstream of steep stepped spill-
Bollaert, E., and A. J. Schleiss. 2003. “Scour of rock due to the impact of ways.”In Proc., 3rd Int. Junior Researcher and Engineering Workshop
plunging high velocity jets. Part I: A state-of-the-art review.” J. Hydraul. on Hydraulic Structures, edited by R. Janssen and H. Chanson.
Res. 41 (5): 451–464. https://doi.org/10.1080/00221680309499991. Brisbane, Australia: Univ. of Queensland.
Bung, D. B., Q. Sun, I. Meireles, T. Viseu, and J. Matos. 2012. “USBR type Pope, S. B. 2000. Turbulent flows. Cambridge, UK: Cambridge University
III stilling basin performance for steep stepped spillways.” In Proc. 4th Press.
IAHR Int. Symp. on Hydraulic Structures. Madrid, Spain: International
Valero, D., and D. B. Bung. 2015. “Hybrid investigations of air transport
Association for Hydro-Environmental Engineering and Research.
processes in moderately sloped stepped spillway flows.” In E-proc.,
Castro-Orgaz, O., and W. H. Hager. 2017. Non-hydrostatic free surface
36th IAHR World Congress. Madrid, Spain: International Association
flows: Advances in geophysical and environmental mechanics and
mathematics. Cham, Switzerland: Springer. for Hydro-Environmental Engineering and Research. https://www.iahr
.org/Portal/About_US/IAHR_World_Congresses/36th_IAHR_World
Downloaded from ascelibrary.org by 105.106.9.188 on 12/12/19. Copyright ASCE. For personal use only; all rights reserved.

Frizell, K. W., and C. D. Svoboda. 2012. Performance of type III stilling


basins: Stepped spillway studies. Do stepped spillways affect tradi- _Congress.aspx?hkey=ad94a6b7-6ad3-4082-8289-b701ac858c92.
tional design parameters? Denver: US Department of Interior, Bureau Valero, D., and R. García-Bartual. 2016. “Calibration of an air entrainment
of Reclamation. model for CFD spillway applications.” In Advances in hydroinfor-
Hunt, S. L., and K. C. Kadavy. 2017. “Estimated splash and training wall matics. Singapore: Springer.
height requirements for stepped chutes applied to embankment dams.” Viti, N., D. Valero, and C. Gualtieri. 2019. “Numerical simulation of
J. Hydraul. Eng. 143 (11): 06017018. https://doi.org/10.1061/(ASCE) hydraulic jumps. Part 2: Recent results and future outlook.” Water
HY.1943-7900.0001373. 11 (1): 28. https://doi.org/10.3390/w11010028.

© ASCE 07019015-3 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(2): 07019015

You might also like