Spillway Jet Regime and Total Dissolved Gas Prediction With A Multiphase Ow Model

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/324643239

Spillway jet regime and total dissolved gas prediction with a multiphase flow
model

Article  in  Journal of Hydraulic Research · April 2018


DOI: 10.1080/00221686.2018.1428231

CITATIONS READS

3 292

3 authors:

Yushi Wang Marcela Politano


University of Iowa University of Iowa
17 PUBLICATIONS   28 CITATIONS    49 PUBLICATIONS   439 CITATIONS   

SEE PROFILE SEE PROFILE

Larry Weber
University of Iowa
104 PUBLICATIONS   1,676 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Iowa Watershed Approach View project

Determining the Minimum Ecological River Inflow Needs of Louisiana Estuaries View project

All content following this page was uploaded by Yushi Wang on 23 April 2018.

The user has requested enhancement of the downloaded file.


Journal of Hydraulic Research

ISSN: 0022-1686 (Print) 1814-2079 (Online) Journal homepage: http://www.tandfonline.com/loi/tjhr20

Spillway jet regime and total dissolved gas


prediction with a multiphase flow model

Yushi Wang, Marcela Politano & Larry Weber

To cite this article: Yushi Wang, Marcela Politano & Larry Weber (2018): Spillway jet regime and
total dissolved gas prediction with a multiphase flow model, Journal of Hydraulic Research

To link to this article: https://doi.org/10.1080/00221686.2018.1428231

View supplementary material

Published online: 18 Apr 2018.

Submit your article to this journal

View related articles

View Crossmark data

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


http://www.tandfonline.com/action/journalInformation?journalCode=tjhr20
Journal of Hydraulic Research, 2018
https://doi.org/10.1080/00221686.2018.1428231
© 2018 International Association for Hydro-Environment Engineering and Research

Research paper

Spillway jet regime and total dissolved gas prediction with a multiphase flow model
YUSHI WANG (IAHR Member), PhD Student, IIHR-Hydroscience & Engineering, 100 C. Maxwell Stanley Hydraulics Laboratory,
The University of Iowa, Iowa City, IA 52242-1585, USA
Email: yushi-wang@uiowa.edu

MARCELA POLITANO (IAHR Member), Research Engineer, IIHR-Hydroscience & Engineering, 100 C. Maxwell Stanley
Hydraulics Laboratory, The University of Iowa, Iowa City, IA 52242-1585, USA
Email: marcela-politano@uiowa.edu (author for correspondence)

LARRY WEBER (IAHR Member), Director, IIHR-Hydroscience & Engineering, 100 C. Maxwell Stanley Hydraulics Laboratory,
The University of Iowa, Iowa City, IA 52242-1585, USA
Email: larry-weber@uiowa.edu

ABSTRACT
A numerical model, based on the open source code OpenFOAM, was developed to predict jet regimes and total dissolved gas downstream of
spillways. The model utilizes the volume of fluid method to track the interface between air and water. A detached eddy simulation model is used
for turbulence closure. Transport and dissolution of bubbles are predicted using an Eulerian approach. A bubble number density equation was
implemented to predict bubble size changes caused by dissolution and compression. Total dissolved gas was computed using a transport equation
that includes the mass transfer between bubbles and water. The model simultaneously captured spillway jet regimes and distribution of total dissolved
gas in a spillway sectional model of McNary Dam. Model parameters, gas volume fraction and bubble size at the entrainment region, were calibrated
to match total dissolved gas measured in the field under different dam operations.

Keywords: Aerated flow; detached eddy simulation model; hydropower; jet regime; OpenFOAM; total dissolved gas; water quality

1 Introduction transport to depth and the production of TDG (Fig. 1). Deflec-
tors are commonly designed to operate in a skimming regime
Total dissolved gas (TDG) can cause gas-bubble disease, which, under a wide range of spillway flowrates and tailwater eleva-
depending on supersaturation level and exposure time, can cause tions. Distinct flow conditions may, however, occur depending
stress or even mortality in affected fish (Elston, 1997; Tan, on the spillway flowrate, tailwater elevation, and deflector char-
2006). Gas supersaturation can occur either as a natural phe- acteristics such as length, curvature radius, and location in the
nomenon or due to hydraulic structures. In hydropower dams, spillway face (Dierking, 2001). At low tailwater elevation, water
bubbles can be entrained along the spillway face and during the flowing over the spillway plunges deep into the stilling basin
plunging of spillway jets in the tailrace pool. TDG supersatu- resulting in a plunging regime with the highest TDG production.
ration occurs if entrained bubbles are transported to deep high On the other hand, a surface jump occurs during high tailwa-
pressure regions in the tailrace where dissolution is enhanced ter elevation. A hydraulic roller forms above the jet, aerating
(Geldert, Gulliver, & Wilhelms, 1998). the downstream water surface with potential of elevated TDG
Hydropower utilities in the Columbia and Snake River production. Sometimes, a transitional regime between surface
Basins in the USA have conducted several laboratory and field jump and skimming flow, known as undular jet, is observed.
investigations to support TDG management and meet water In this regime, the spillway jet ramps up on the downstream
quality standards to protect fish. One alternative to reduce TDG water surface and a flow recirculation below the jet starts to
supersaturation is to install spillway deflectors to change the develop. Deflector performance is usually tested in a reduced-
regular plunging jet into skimming flow minimizing the bubble scale laboratory model before installation in the field (Nielsen,

Received 13 February 2017; accepted 11 January 2018/Currently open for discussion.

ISSN 0022-1686 print/ISSN 1814-2079 online


http://www.tandfonline.com
1
2 Y. Wang et al. Journal of Hydraulic Research (2018)

Figure 1 Spillway jet regimes: (a) plunging flow, (b) skimming flow, (c) undular jet and (d) surface jump

Weber, & Haug, 2000). This approach has demonstrated to be the spillway using the volume of fluid (VOF) method. The free-
useful to infer TDG production through the observed spillway surface shape was then fixed during the computation of the
regimes. However, since bubbles cannot be scaled in the lab- hydrodynamics, bubble and TDG in the entire tailrace. This
oratory, TDG generation cannot be fully evaluated with this approach facilitated the implementation of attenuation of normal
methodology. Furthermore, deflectors possibly perform differ- fluctuations at the free surface and reduced required compu-
ently in the field since bubbles and turbulence, which are not tational resources. Quantitative agreement between numerical
properly represented in the laboratory model, play an important results and TDG field data was obtained at several dams with
role in the resulting jet regime. this model (Arenas Amado, Politano, & Anderson, 2011; Poli-
Early attempts to predict TDG downstream of dams were tano, Arenas Amado, et al., 2009; Politano, Carrica, & Weber,
based on laboratory and field experiments and data fitting 2009; Politano, Arenas Amado, Bickford, Murauskas, & Hay,
(Geldert et al., 1998; Roesner & Norton, 1971). One draw- 2011; Turan, Politano, Carrica, & Weber, 2007). However, use
back of this approach is that the derived empirical correlations of a fixed free surface prevented the use of the model to unsteady
for TDG are limited to the geometry and range of operations conditions.
used to obtain the model parameters. Orlins and Gulliver (2000) Recent development of faster computers allows the use of
developed a two-dimensional (2D), laterally-averaged numer- mechanistic models that, for the first time, can simultaneously
ical model to predict TDG generation and transport. Their predict spillway jet regimes, as a function of spillway flowrate
model used a TDG transport equation, accounting for con- and tailwater elevation, and TDG distribution. This work is a
vection, turbulent diffusion, and mass transfer across the free follow-up of the model developed by Wang, Politano, Laughery,
surface. Hydrodynamic data including flow velocity, turbulent and Weber (2015) using a VOF model with detached eddy simu-
variables, and bubble distribution were obtained from a reduced- lation (DES) to predict hydrodynamics and spillway jet regimes
scale laboratory model. TDG field measurements were used to under varying flow conditions. Good agreement between obser-
determine model parameters. Weber, Huang, Lai, and McCoy vations in the reduced-scale laboratory model and numerical
(2004) improved the model by predicting the hydrodynamics simulations was found. In the current paper, the authors cou-
with a rigid-lid computational fluid dynamics (CFD) model and pled the VOF-DES model with an Eulerian two-phase flow
extending the model to three dimensions. The rigid-lid model model using the OpenFOAM open-source library. The advan-
was not intended for spillway jet regimes prediction. Urban, tage of this model is the ability to capture the observed free
Gulliver, and Johnson (2008) developed governing equations for surface unsteadiness near the spillway and thus better represent
three distinct flow regions downstream of a spillway. The flow the whole physics of the TDG phenomenon near the dam.
field was calculated following a relationship developed in an
experimental study by Ead and Rajaratnam (2002). Note that in
the models mentioned above, bubble distribution in the tailrace 2 Mathematical model
was not fully simulated, but was estimated based on empirical
correlations. To predict tailrace hydrodynamics, bubble trans- The appropriate modelling of the free surface and turbulence is
port and TDG distribution, Politano, Carrica, Turan, and Weber important to predict spillway jet regimes and TDG distribution.
(2007) used a 2D two-phase flow model. A bubble number den- Isotropic Reynolds-averaged Navier–Stokes models (RANS),
sity transport equation was introduced to account for bubble commonly used in hydraulic applications, overpredict turbu-
size variation due to mass transfer and compression. A rigid-lid lence levels near the free surface, resulting in weaker spillway
non-flat approach was used for TDG calculation. Their model jets than observed (Turan et al., 2007) and ineffectual transport
focused on the bubble–liquid mass transfer and TDG transport of bubbles and TDG production. A DES approach, which blends
in the tailrace, assuming a known entrained gas volume fraction RANS near walls with large eddy simulation (LES) in the main
and bubble size distribution. In a later work, Politano, Arenas flow, can capture the anisotropic behaviour of the turbulence
Amado, and Weber (2009) predicted the free-surface shape near near the free surface, reducing the computational time required
Journal of Hydraulic Research (2018) Modelling spillway jet regimes and TDG 3

by LES simulations. Model details of a free surface model to The bubble diameter is calculated from:
simulate spillway jets at reduced and prototype scales are found
  13
in Wang et al. (2015). 6α
Db = (7)
Witt, Gulliver, and Shen (2015) used the VOF method to πN
reproduce air entrainment in a laboratory-scale hydraulic jump.
Since grid size needs to be of the order of the bubble diameter, The total dissolved gas concentration C is computed from:
using this approach to represent bubbles in a full-scale sectional
∂C  νt

model is beyond the most powerful computers currently avail- + ∇ · (UC) = ∇ · ν + ∇C + S (8)
able. To overcome this limitation, the Eulerian model proposed ∂t Sc
by Politano et al. (2007) was used in this study. As a first step,
where U is the fluid velocity vector. ν and νt are the fluid kine-
the model uses a one-way coupling approach assuming a neg-
matic molecular and turbulent viscosity, respectively, and Sc is
ligible influence of the bubbles on the liquid field. Continuity
the Schmidt number. The rate of mass transfer between bubble
and momentum equations for the bubble phase are (Politano,
and liquid phases can be modelled as (Politano, Fu, et al., 2009):
Fu, Luellen, & Weber, 2009):
 
∂α p + (2 σ/Db )
+ ∇ · (Ub α) = −S (1) S= π NDb2 kl −C (9)
∂t kh

where kl is the mass transfer coefficient due to turbulence, σ is


0 = −α∇p + αρb g + M b (2) the interfacial tension, and kh is Henry’s constant.
In this study, the bubble velocity computed with Eq. (2)
where α is the gas volume fraction, Ub is bubble velocity vector,
results in bubbles free to flow across the free surface. Mass
p is the pressure, t is time, g is the gravitational field, and ρb is
transfer at the free surface was neglected in this study since
bubble density calculated following the ideal gas law:
in most conditions bubble dissolution is the dominant process
Mp in the aerated zone near the spillway (Urban et al., 2008).
ρb = (3) Future work will include the surface mass transfer, which can
RT
be important at highly supersaturated conditions.
with M the average molar mass of air, R the universal gas con-
stant, and T the temperature. S represents the bubble–liquid
3 Site description
mass transfer and M b is the interfacial momentum transfer
between phases. In this study, it is assumed that drag is the dom-
Model results were compared against TDG measurements
inant interfacial force downstream of spillways. Future work
downstream of McNary Dam (Wilhelms, Carroll, & Schneider,
involves the inclusion of turbulent dispersion and virtual mass
1997). McNary is located in the Columbia River between Ore-
to better predict the bubble distribution in the tailrace. The
gon and Washington and operated by the US Army Corps of
interfacial momentum transfer vector is obtained from:
Engineers, Walla Walla District. The dam, one of the largest
3 CD hydroelectric power facilities in the Pacific Northwest, is a
M b = − ρl α Ur |Ur | (4) 2.2 km long concrete gravity dam with 22 vertical spillway
4 Db
gates, each 15.24 m by 15.54 m (Fig. 2). Spillbay 1 is on the
with Ur the relative velocity vector of the bubble with respect to north side of the river and close to the fish ladder. Spillbay 22 is
the liquid phase, ρl the liquid density and Db the bubble diame- on the south side and close to the powerhouse. To reduce TDG
ter. The drag coefficient CD can be modelled as (Ishii & Zuber, in the tailrace of the dam, spillway deflectors were designed in a
1979; Tomiyama, 1998): 1:25 scale laboratory model that included one full central spill-
bay and two half spillbays (Northwest Hydraulic Consultants


⎪ 24(1 + 0.15 R0.687 ) [NHC], 2001).
⎨ b
for Db ≥ 2e−4
R Deflectors were installed in McNary Dam first in spillbays
CD = b (5)

⎪ 24 3 through 20 and TDG measurements were collected directly
⎩ for Db < 2e−4
Rb below the spillway on a grid of instruments set on five lateral
transects and three longitudinal profiles (Wilhelms et al., 1997).
where Rb = ρl Db |Ur |/μl is the bubble Reynolds number, with The instruments were mounted along five steel cables anchored
μl the dynamic viscosity. The transport equation for bubble to the spillway piers between spillbays 1 and 2, spillbays 4 and
number density N is (Politano, Fu, et al., 2009): 5, spillbays 11 and 12, spillbays 17 and 18, and spillbays 20
and 21. Transects were located approximately 120, 190 and 200
∂N m, downstream of the piers. Two instruments were installed at
+ ∇ · (Ub N ) = 0 (6)
∂t most locations: a fixed instrument at the bottom and a floating
4 Y. Wang et al. Journal of Hydraulic Research (2018)

Figure 2 McNary Dam

instrument that provided surface-to-mid-depth measurements. saturation for each flow condition, only measurements in the
The bottom and surface instruments measured TDG saturation second hour of each spill pattern were used to compare against
at 15 and 5 min intervals, respectively. Cables were initially model results.
installed perpendicular to the spillways; however, the position
of the surface-to-mid-depth instruments changed throughout the
testing period in response to drag on the instruments, cables, and
4 Numerical model
buoys. TDG concentrations collected by the bottom instruments
were used to compare with model results since vertical position-
4.1 Code and numerical method
ing of the surface-to-mid-depth instruments was not available
(Wilhelms et al., 1997). During the field study, each spill pattern The solver interFOAM was used to solve the two incom-
was set for a minimum duration of two hours. pressible immiscible fluids (water–air) using a VOF interface
capturing approach. The PIMPLE algorithm (a merged Pressure
Implicit with Splitting of Operators (PISO) and Semi-Implicit
3.1 Operational conditions
Method for Pressure Linked Equations (SIMPLE) algorithm)
Four operational conditions observed during the field study were was used to couple pressure and velocity. The first-order implicit
used to evaluate the capability of the model to predict spillway Euler method was applied to the unsteady terms. In a DES simu-
jet regimes and TDG distribution (Table 1). Conditions were lation, central schemes are generally preferred. In this study, the
chosen based on the deflector performance curve obtained in limited linear differencing scheme was used as the discretization
NHC (2001). Spillway flow regimes were mostly defined by method for the spatial derivative terms to balance accuracy and
the spillway discharge since tailwater elevation varied over a stability.
small range between 80.8 m to 81.9 m. TDG measurements Equations (1)–(9) were incorporated into the standard inter-
showed a transient period of about 30 min before a quasi-steady FOAM solver to predict the bubble distribution and bubble size.
state condition was established. To ensure a representative TDG A compression term analogous to that of the phase indicator
function, ∇ · (γ (1 − γ )Ur ), where γ is the liquid volume frac-
Table 1 Simulation conditions tion, was included on the left-hand side of Eq. (8) to avoid
diffusion at the interface (Bohorquez, 2008). The specialized
Discharge (m3 s−1 ) Water surface elevation (m) scheme proposed by Rusche (2003) was applied for this term.
Spillbays 4 and 5 Forebay Tailwater Flow regime
Unsteady solutions were obtained using a variable time step
with a fixed Courant number of 0.4. Zero velocity and turbu-
385.1 103.2 81.4 Plunging lence were imposed as initial conditions for the entire domain.
201.0 103.1 80.8 Skimming The gas volume fraction was assumed zero and measured
155.7 103.0 81.9 Undular forebay TDG was imposed over the entire tailrace as initial
107.6 103.0 81.5 Hydraulic jump
condition.
Journal of Hydraulic Research (2018) Modelling spillway jet regimes and TDG 5

4.2 Model domain and computational grid flows under sluice gates. Turbulent variables are assumed
to be zero.
In the present study, only part of the McNary Dam tailrace
(2) Outflow: The boundary at the downstream end is split at
was simulated to reduce the computational resources required.
the tailwater elevation into two domains that allow only
Rigid-lid CFD simulations of the McNary tailrace under diverse
one fluid to flow through. Hydrostatic pressure is used
dam configurations indicated minimal effect of lateral flows
for the water below the tailwater elevation. Atmospheric
downstream of spillbays on the north side of the dam (Poli-
pressure is imposed for air. Zero-gradient is used for all
tano & Dvorak, 2012). Since spillbays 1 and 2 use a special
other variables.
split-leaf gate that affect normal air entrainment and spillway
(3) Walls and river bed: A no-slip condition and zero flux for all
regimes, measurements downstream of spillbays 4 and 5 were
other variables are used at the river bed and walls. Follow-
used in this study. The model included two half spillbays, flow
ing Meselhe and Odgaard (1998), the effect of roughness
deflectors, sluice gate and approximately 380 m of the tailrace
was neglected in this study.
using measured bathymetry (Fig. 3a).
(4) Symmetry: Symmetry conditions with zero shear stress and
Numerical grids, containing approximately 0.5 million cells,
zero flux for all quantities were used at adjacent spillbays.
were based on the grid sensitivity study by Wang et al. (2015).
(5) Top: A pressure outlet boundary condition with atmospheric
Grids were refined near the expected free surface to avoid
pressure is applied at the top of the computational domain
numerical diffusion (Fig. 3a–e).
to allow free flow of air and avert unrealistic pressure
values.
4.3 Boundary conditions

Boundary conditions used in the model are shown in Fig. 3a and 4.4 Model parameters
summarized below: Predicting the air entrainment with a free surface tracking
method would require a grid size small enough to resolve the
(1) Inflow: water depth and velocity at the contraction down- entrained bubbles, which is beyond current computer capabil-
stream of the spillway gates are imposed at the upstream ities (Castro, Li, & Carrica, 2016). To the best knowledge of
end following the approach used by Kim (2007) to study the authors, measurements of entrained air are not available

Figure 3 Numerical domain and boundary conditions: (a) 3D mesh, (b) 2D view of mesh for plunging flow, (c) 2D view of mesh for skimming
flow, (d) 2D view of mesh for undular jet and (e) 2D view of mesh for surface jump
6 Y. Wang et al. Journal of Hydraulic Research (2018)

for any spillway at prototype scale. Castro et al. (2016) pro- Yang (2016) assumed a normal distribution of bubble size.
posed a model for air entrained in free surface turbulent flows. A simplified approach assuming a constant gas volume frac-
The authors calibrated their model against breaking waves at tion and bubble size at the model upstream end is used in
laboratory scale and measurements in a full-scale Navy ship. this study to perform a sensitivity analysis of these parameters.
Implementation of Castro’s entrainment model is proposed After entrainment, the model takes into account that gas volume
as future research, after two-phase flow data downstream of fraction and bubble size change due to dissolution and pres-
spillways become available to support the calibration of the sure. Simulations were performed using at the model upstream
model. end: (1) constant bubble diameter of 0.8 mm and gas volume
In this study, the amount and size of bubbles entrained in fractions ranging from α = 0.01 to α = 0.03; and (2) con-
the spillway are model parameters. Bubble diameters entrained stant gas volume fraction of 0.03 with bubble diameters rang-
as a result of plunging liquid jets have a bimodal distribu- ing from Db = 0.1 to Db = 2mm. These values were selected
tion with fine primary bubbles smaller than 1 mm and larger based on data used to match field TDG data with numerical
secondary bubbles of approximately 4 mm (Biń, 1993; Wood, predictions with a rigid-lid 3D model in several hydropower
1991). Small bubbles are the result of the distribution produced dams (Politano, Arenas Amado, et al., 2009; Politano, Car-
by the breakup of the large bubbles. 2D numerical simula- rica, et al., 2009; Politano, Fu, et al., 2009; Politano et al.,
tions by Politano et al. (2007) and Ma, Li, Hodges, Feng, and 2011).

Figure 4 Evolution of TDG at the downstream end: (a) plunging flow, (b) skimming flow, (c) undular jet and (d) surface jump

Figure 5 Velocity vectors and gas volume fraction contours: (a) plunging flow, (b) skimming flow, (c) undular jet and (d) surface jump
Journal of Hydraulic Research (2018) Modelling spillway jet regimes and TDG 7

Figure 6 TDG and bubble diameter contours: (a) plunging flow, (b) skimming flow, (c) undular jet and (d) surface jump

5 Model results formed above the deflector in the surface jump condition. These
bubbles leave the free surface faster resulting in less gas vol-
5.1 Spillway flow regime, bubble distribution and TDG ume fraction in the downstream channel for the surface jump
concentration than for the undular condition. This trend was also observed in a
bubble entrainment study in a reduced-scale spillway (Hoschek,
Simulations using an entrained gas volume fraction of 3% and
Carrica, & Weber, 2008).
bubble diameter of 0.8 mm were carried out to evaluate the
Figure 6 shows bubble diameter as contour lines and contours
general flow pattern and TDG production/degasification under
of TDG concentration, for all flow regimes. Bubbles shrunk at
different flow regimes. Figure 4 shows the evolution of the
depth due to compression and dissolution. On the other hand,
area-weighted averaged TDG at the model downstream end for
bubbles absorbed air and become larger near the surface under
all simulations. TDG concentration gradually developed until
low pressure and supersaturated condition. Bubble size has a
reached a quasi-steady state condition at about 300 s to 1000 s,
strong effect on the dissolution since smaller bubbles have a
depending on the flowrate. Since the tailwater elevation only
longer residence time (less buoyancy force) and a larger interfa-
varied in a small range, increasing the flowrate resulted in larger
cial area per unit volume. The TDG distribution is determined
velocities in the downstream channel and faster convergence.
by convective transport, bubble dissolution, and turbulent mix-
Figures 5–9 show time-averaged variables at 1500 s. Gas vol-
ing. Dissolution was important between the deflector and the
ume fraction contours and velocity vectors in Fig. 5 show the
first row of baffle piers, where bubbles were transported to
distinctive flow pattern and surface waves predicted for dif-
depth. The highest TDG concentration was found downstream
ferent spillway flowrates. The animation of instantaneous free
of the end sill for the plunging flow condition, reaching about
surface location and gas volume fraction in the online sup-
165%. The highest TDG concentration for the skimming flow,
plemental data of the paper (https://doi.org/10.1080/00221686.
undular jet and surface jump conditions were approximately
2018.1428231) show the highly dynamic nature of the flow
143%, 153%, and 153%, respectively. The dissolution decreased
downstream of spillways. Flow regimes are consistent with
near the end sill, though remained high compared to the down-
the deflector performance curve obtained in the reduced-scale
stream channel. TDG levels gradually decreased as they move
laboratory model (NHC, 2001). The model captured the general
away from the spillway due to the degasification near the free
bubble behaviour observed in laboratory models (Babb, Christi-
surface and turbulent mixing. Bubbles near the free surface
son, Cain, & Ahmann, 2000; Cain, 1997; Dierking, 2001; Weber
promoted degasification when TDG was larger than that at
& Mannheim, 1997). Plunging jets at high spillway flowrate
equilibrium at the local bubble depth.
entrained bubbles to depth increasing the gas volume fraction
in the tailrace. In contrast, the skimming flow observed at low
5.2 Comparison with field data and effect of model
spillway flowrate greatly reduced the downward bubble trans-
parameters on TDG distribution
port. At medium spillway flowrates, undular and surface jump
conditions were predicted as observed in the laboratory. In these Predicted TDG concentration with standard deviation along the
regimes, the spillway jet is first deflected upward and then centre line of the spillbays 4 and 5, for different gas volume
splashes down with the remaining downward momentum. The fractions and bubble diameters, are plotted in Figs 7 and 8,
surface jump dissipated the spillway jet in a shorter distance. A respectively. TDG measurements are also shown with symbols
strong vertical recirculation was predicted beneath the spillway in the same figures. In Fig. 9, predicted TDG concentration as a
jet for the plunging and surface jump conditions. Some bubbles function of the model parameters is shown for different spillway
were trapped by this recirculation until they were completely regimes at the upstream (Q4T1B) and downstream (Q4T2B)
dissolved. Some bubbles were also trapped in the roll-over wave TDG sensor locations.
8 Y. Wang et al. Journal of Hydraulic Research (2018)

Figure 7 Effect of entrained gas volume fraction on TDG distribution: (a) plunging flow, (b) skimming flow, (c) undular jet and (d) surface jump
Journal of Hydraulic Research (2018) Modelling spillway jet regimes and TDG 9

Figure 8 Effect of entrained bubble size on TDG distribution: (a) plunging flow, (b) skimming flow, (c) undular jet and (d) surface jump

Gas dissolution and resulting TDG increased with the gas volume fraction, fewer bubbles were available in the water
entrained gas volume fraction. The standard deviation is much column, resulting in lower TDG production in the stilling basin
higher in the stilling basin than in the downstream channel and and minor degasification along the channel. In this case, pre-
this effect is stronger for larger entrained gas volume fraction. dicted TDG levels in the upstream and downstream stations
The TDG concentration gradually decreased downstream of the were similar. The degasification rate was slightly higher for
stilling basin due to the degasification by bubbles near the free larger entrained gas volume fraction due to two main mech-
surface and turbulent mixing within the water column. For low anisms: more bubbles were available near the free surface to
10 Y. Wang et al. Journal of Hydraulic Research (2018)

Figure 9 Predicted TDG concentration as a function of gas volume fraction (left) and bubble diameter (right): (a) and (b) plunging flow, (c) and (d)
skimming flow, (e) and (f) undular jet, and (g) and (h) surface jump

absorb air from the liquid and TDG concentration reduced to and TDG production. For all spillway regimes, predicted TDG
equilibrium faster at higher gas supersaturation. concentration decreased with the size of entrained bubbles.
The effect of bubble size on TDG distribution is less straight- Due to buoyancy, larger bubbles with higher terminal velocities
forward than the gas volume fraction. The size of entrained move closer to the free surface, minimizing TDG production.
bubble affects the distribution of bubbles in the water column In addition, the interfacial area of larger bubbles is small,
Journal of Hydraulic Research (2018) Modelling spillway jet regimes and TDG 11

Table 2 Comparison between predicted and measured TDG The present work is the initial step towards the development
of a more advanced tool for TDG prediction. The gas volume
Predicted TDG Measured mean TDG
fraction and bubble size distribution at the entrainment region
Station Averaged Sn Averaged Sn Deviation (%) are the experimental model parameters. A sensitivity analysis
of these parameters indicates that significant research effort is
Plunging
needed to obtain measurements of two-phase flow data at pro-
Q4T1B 1.4671 0.0354 1.4737 0.0090 − 0.45
Q4T2B 1.4225 0.0202 1.3779 0.0040 3.23 totype scale. After these data become available, a model for air
Skimming entrainment can be implemented and calibrated to improve the
Q4T1B 1.2726 0.0136 1.2775 0.0033 − 0.38 current model. In addition, a polydisperse two-phase flow model
Q4T2B 1.2513 0.0057 1.2537 0.0019 − 0.19 that accounts for breakup and coalescence is proposed as future
Undular
work to predict the evolution of the bubble size distribution after
Q4T1B 1.2867 0.0103 1.2918 0.0024 − 0.39
Q4T2B 1.2656 0.0066 1.2633 0.0017 0.19 entrainment. Bubble turbulent dispersion and degasification at
Hydraulic jump the free surface can be included to further improve the model.
Q4T1B 1.2109 0.0064 1.2285 0.0027 1.43 However, constitutive equations for the high Reynolds numbers
Q4T2B 1.1960 0.0019 1.1845 0.0022 − 0.97 present in prototype scale are inexistent, and the applicability
of these models downstream of spillways requires, again, two-
phase field measurements for calibration. It is expected that the
further reducing the dissolution rate. According to the model, development of faster and more accessible computers in the next
bubbles with 2 mm diameter or larger did not significantly con- few years together with a set of two-phase flow field data will
tribute to TDG production in the tailrace. When the bubbles allow a complete model validation and use of the model for fully
are sufficiently small to be carried to the river bottom, TDG evaluation of TDG mitigation measures.
concentration near the river bed depended on the river bed ele-
vation. Note that for 0.1 mm entrained bubbles, the predicted
TDG at the upstream sensor was lower than downstream due to Funding
the higher river bed elevation at the upstream location.
Field observations and laboratory experiments suggest that This work was supported by the Hydro Research Foundation
the entrained bubble gas volume fraction can vary significantly [grant number DE-EE0002668].
under different flow regimes (Hoschek et al., 2008). However,
entrained bubble size that significantly contributes to TDG is
Supplemental data
usually within a relatively small range for varying conditions.
With the assumed entrained bubble diameter of 0.8 mm, the gas
Supplemental data for this article can be accessed doi:10.1080/
volume fraction values that better match the measured TDG
00221686.2018.1428231
under flow regimes plunging, skimming, undular, and surface
jump are 0.02, 0.0225, 0.0125, and 0.0085, respectively.
Predicted TDG values after calibration at each sensor loca- Notation
tion along with field measurements and standard deviation, Sn ,
are summarized in Table 2. Differences between predicted and C = TDG concentration (–)
measured TDG values are within 4%. Note that the reduction Db = bubble diameter (m)
in TDG measured in the longitudinal direction is higher than g = gravitational field (m s−2 )
predicted with the model for the plunging and surface jump con- kh = Henry’s constant (N m−2 )
ditions. This is likely due to the dilution with the lateral flow kl = mass transfer coefficient due to turbulence (–)
from the powerhouse region, which was not captured by the M = average molar mass of air (g mol−1 )
present sectional spillway model. N = bubble number density (–)
p = pressure (Pa)
6 Conclusions and future work Q = flowrate at the gate (m3 s−1 )
R = universal gas constant (J mol−1 k−1 )
A mechanistic multiphase flow model based on the open-source Rb = bubble Reynolds number (–)
OpenFOAM was developed to simultaneously predict, for the Sc = Schmidt number (–)
first time, spillway jet regimes and TDG distribution. TDG is t = time (s)
calculated with a two-phase transport equation in which the T = temperature (k)
source is the bubble dissolution, function of the gas volume frac- U = liquid velocity vector (m s−1 )
tion and bubble size. Model results were compared against TDG Ub = bubble velocity vector (m s−1 )
field data collected in McNary Dam. The model was able to cap- Ur = relative velocity vector of the bubble with respect to the
ture the observed spillway jet regimes and longitudinal TDG liquid phase (m s−1 )
distribution at different dam operations. α = gas volume fraction of bubbles (–)
12 Y. Wang et al. Journal of Hydraulic Research (2018)

γ = liquid volume fraction for the interface water/air (–) Geldert, D. A., Gulliver, J. S., & Wilhelms, S. C. (1998). Mod-
μl = liquid dynamic viscosity (kg m−1 s−1 ) eling dissolved gas supersaturation below spillway plunge
ν = liquid kinematic molecular viscosity (m2 s−1 ) pools. Journal of Hydraulic Engineering, 124(5), 513–521.
νt = liquid kinematic turbulent viscosity (m2 s−1 ) Hoschek, S. S., Carrica, P. M., & Weber, L. J. (2008). Bub-
ρl = liquid density (kg m−3 ) ble entrainment and distribution in a model spillway with
ρb = bubble density (kg m−3 ) application to total dissolved gas minimization. Journal of
σ = interfacial tension (N m−1 ) Hydraulic Engineering, 134(6), 763–771.
Ishii, M., & Zuber, N. (1979). Drag coefficient and relative
velocity in bubbly, droplet or particulate flows. AIChE Jour-
nal, 25(5), 843–855.
References Kim, D.-G. (2007). Numerical analysis of free flow past a sluice
gate. KSCE Journal of Civil Engineering, 11(2), 127–132.
Arenas Amado, A., Politano, M., & Anderson, K. (2011). A Ma, Q., Li, R., Hodges, B., Feng, J., & Yang, H. (2016). Two-
CFD model to evaluate the effect of sluiceway deflectors on phase flow simulation of supersaturated total dissolved gas
the hydrodynamics and TDG field in the tailrace of Hells in the plunge pool of a high dam. Journal of Hydraulic
Canyon Dam. In E. M. Valentine, C. J. Apelt, J. Ball, H. Engineering, 124(12), 1203–1214.
Chanson, R. Cox, R. Ettema, . . . J. E. Sargison (Eds.), Pro- Meselhe, E. A., & Odgaard, A. J. (1998). 3D numerical flow
ceedings of the 34th world congress of the international model for fish diversion studies at Wanapum Dam. Environ-
association for hydro-environment research and engineer- mental Progress & Sustainable Energy, 35(4), 1139–1148.
ing: 33rd hydrology and water resources symposium and Nielsen, K. D., Weber, L., & Haug, P. E. (2000). Hydraulic
10th conference on hydraulics in water engineering (pp. model study for fish diversion at Wanapum/Priest Rapids
3183–3190). Barton, A.C.T.: Engineers Australia. development, Part XVI: 1:32.5 scale sectional model of
Babb, A. F., Christison, K., Cain, J. D., & Ahmann, M. (2000). Wanapum Dam spillway deflectors (Report No. 284). IIHR
Spillway deflectors for dissolved gas reduction at McNary Hydroscience & Engineering, The University of Iowa.
Dam. In U. Maione, B. Majone-Lehto, & R. Monti (Eds.), Northwest Hydraulic Consultants (NHC). (2001). McNary Dam
Proceedings of the new trends in water and environmen- spillway flow deflectors hydraulic model study – Final report.
tal engineering for safety and life (pp. 85–95). Balkema, Seattle, WA: Author.
Rotterdam: CRC Press. Orlins, J. J., & Gulliver, J. S. (2000). Dissolved gas supersat-
Biń, A. K. (1993). Gas entrainment by plunging liquid jets. uration downstream of a spillway II: Computational model.
Chemical Engineering Science, 48(21), 3585–3630. Journal of Hydraulic Research, 38(2), 151–159.
Bohorquez, P. (2008). Study and numerical simulation of sed- Politano, M., Arenas Amado, A., Bickford, S., Murauskas,
iment transport in free-surface flow (Ph.D. dissertation). J., & Hay, D. (2011). Investigation into the total dissolved
University of Malaga. Retrieved from https://infoscience.epfl. gas dynamics of Wells Dam using a two-phase flow model.
ch/record/130534/files/thesisBohorquezPatricio.pdf Journal of Hydraulic Engineering, 137(10), 1257–1268.
Cain, J. D. (1997). Design of spillway deflectors for ice harbor Politano, M., Arenas Amado, A., & Weber, L. (2009). An inves-
dam to reduce supersaturated dissolved gas levels down- tigation into the total dissolved gas dynamics of the Wells
stream. In J. Gulliver (Ed.), Energy and water: Sustainable project (Report No. 2149). FERC.
development. Proceeding of the 27th congress of the IAHR Politano, M., Carrica, P., & Weber, L. (2009). A multiphase
(pp. 607–612). San Francisco, CA: ASCE. model for the hydrodynamics and total dissolved gas in
Castro, A. M., Li, J., & Carrica, P. M. (2016). A mecha- tailraces. International Journal of Multiphase Flow, 35(11),
nistic model of bubble entrainment in turbulent free sur- 1036–1050.
face flows. International Journal of Multiphase Flow, 86, Politano, M., & Dvorak, J. (2012). Computational Fluid
35–55. Dynamics (CFD) modeling to support the reduction of
Dierking, P. B. (2001). Hydraulic modeling of Hells Canyon fish passage exposure to predator habitat at McNary Dam
Dam for spillway deflector design (Ph.D. dissertation). The (Report No. 377). IIHR Hydroscience & Engineering. Iowa
University of Iowa. City, IA: The University of Iowa.
Ead, S. A., & Rajaratnam, N. (2002). Plane turbulent wall jets in Politano, M., Fu, X., Luellen, L., & Weber, L. (2009). Compu-
shallow tailwater. Journal of Engineering Mechanics, 128(2), tational fluid dynamics model to evaluate total dissolved gas
143–155. field within the tailrace at McNary Dam (Report No. 359).
Elston, R. (1997). Fish kills in resident and captive fish caused IIHR Hydroscience & Engineering, The University of Iowa.
by spill at Grand Coulee Dam in 1997: final report. Pre- Politano, M. S., Carrica, P. M., Turan, C., & Weber, L. (2007).
pared for the Confederated Tribes of the Colville Reserva- A multidimensional two-phase flow model for the total dis-
tion and Columbia River Fish Farms, Aquatechnics, Omak, solved gas downstream of spillways. Journal of Hydraulic
Washington. Research, 45(2), 165–177.
Journal of Hydraulic Research (2018) Modelling spillway jet regimes and TDG 13

Roesner, L. A., & Norton, W. R. (1971). A Nitrogen gas (N2) Weber, L. J., Huang, H., Lai, Y., & McCoy, A. (2004). Modeling
model for the Lower Columbia River (Final Report). Walnut total dissolved gas production and transport downstream of
Creek (CA): Water Resources Engineers. spillways: Three-dimensional development and applications.
Rusche, H. (2003). Computational fluid dynamics of dispersed International Journal of River Basin Management, 2(3), 157–
two-phase flows at high phase fractions (Doctoral disserta- 167.
tion). Imperial College London (University of London). Weber, L. J., & Mannheim, C. (1997). A unique approach for
Tan, D. C. (2006). Research on the lethal effect of the dissolved physical model studies of nitrogen gas supersaturation. In J.
gas supersaturation resulted from Three Gorges project to Gulliver (Ed.), Energy and water: Sustainable development.
Fish (Ph.D. dissertation). Southwest University (in Chinese). Proceeding of the 27th congress of the IAHR (pp. 518–523).
Tomiyama, A. (1998). Struggle with computational bubble San Francisco, CA: ASCE.
dynamics. Multiphase Science and Technology, 10(4), 369– Wilhelms, S. C., Carroll, J., & Schneider, M. L. (1997). Near-
405. field study of total dissolved gas in the McNary spillway
Turan, C., Politano, M. S., Carrica, P. M., & Weber, L. (2007). tailwater. CEWES-CR-F Memorandum for Record dated 22
Water entrainment due to spillway surface jets. Interna- August 1997, U.S. Army Engineer Waterways Experiment
tional Journal of Computational Fluid Dynamics, 21(3–4), Station.
137–153. Witt, A., Gulliver, J., & Shen, L. (2015). Simulating air entrain-
Urban, A. L., Gulliver, J. S., & Johnson, D. W. (2008). Modeling ment and vortex dynamics in a hydraulic jump. International
total dissolved gas concentration downstream of spillways. Journal of Multiphase Flow, 72, 165–180.
Journal of Hydraulic Engineering, 134(5), 550–561. Wood, I. R. (1991). Air entrainment in free-surface slows. IAHR
Wang, Y., Politano, M., Laughery, R., & Weber, L. (2015). hydraulic structures Design Manual No. 4, Hydraulic Design
Model development in OpenFOAM to predict spillway Considerations, Balkema Publ., Rotterdam, The Netherlands,
jet regimes. Journal of Applied Water Engineering and 149 pages.
Research, 3(2), 80–94.

View publication stats

You might also like