You are on page 1of 11

Proceedings of the ASME 2020

Verification and Validation Symposium


VVS2020
May 20-22, 2020, Virtual, Online

VVS2020-8821

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
CFD VERIFICATION AND VALIDATION EXERCISE: TURBULENT MIXING OF
STRATIFIED LAYER

Jason Thompson, Christopher Boyd


U.S. Nuclear Regulatory Commission, Washington, D.C.
Email: jason.thompson@nrc.gov

ABSTRACT nation’s civilian use of radioactive material to protect public


The US Nuclear Regulatory Commission (NRC) health and safety. These principles are laid out in the NRC’s
participated in an Organization for Economic Cooperation and mission statement as well as in international guidelines such as
Development / Nuclear Energy Agency (OECD/NEA) the Organization for Economic Co-operation and Development
benchmark activity based on testing in the PANDA facility (OECD) Nuclear Energy Agency (NEA) report “The
located at the Paul Scherrer Institute in Switzerland. In this test, Characteristics of an Effective Nuclear Regulator” (NEA 2014,
a stratified helium layer was eroded by a turbulent jet from [1]). The regulatory mission includes the review and evaluation
below. NRC participated in this benchmark to develop expertise of licensee-submitted analyses that support the safety
and modeling guidelines for computational fluid dynamics justification for specific reactor-system components or
(CFD) in anticipation of utilizing these methods for future safety scenarios. In many cases, these analyses involve codes that have
and confirmatory analyses. CFD predictions using ANSYS been approved for the specific application of interest after the
FLUENT V19.0 are benchmarked using the PANDA test data, corresponding approach has developed a proven history of
and sensitivity studies are used to evaluate the significance of validation and acceptance in the regulatory environment. The
key phenomena, such as boundary conditions and modeling application of CFD in regulatory activities is not yet common but
options, that impact the helium erosion rates and jet velocity the number of applications and interest in this area is growing.
distribution. The k-epsilon realizable approach with second Increased computational power along with a growing number of
order differencing resulted in the best prediction of the test data. capable analysts involved with the design and safety analyses of
The most significant phenomena are found to be the inlet mass new reactor systems are factors that will likely lead to continued
flowrate and turbulent Schmidt number. CFD uncertainty for growth in the use of CFD for nuclear regulatory safety
helium and velocity due to numerical error and input parameter applications.
uncertainty are predicted using a sensitivity coefficient The NRC recognizes the growth of CFD methods and is
approach. Numerical uncertainty resulting from the mesh design leveraging resources to further develop and maintain capabilities
is estimated using a Grid Convergence Index (GCI) approach. in this area. This includes developing knowledge and expertise
Meshes of 0.5, 1.5 (base mesh), and 4.5 million cells are used for through cooperative initiatives, benchmark exercises, specific
the GCI. Approximately second order grid convergence was test programs, targeted CFD model developments, verification
observed but p (order of convergence) values from 1 to 5 were and validation initiatives, and the support for and development
common. The final helium predictions with one-sigma of CFD best-practice guidelines. The NRC uses CFD, where
uncertainty interval generally bounded the experimental data. appropriate, to support the resolution of safety issues and for
The predicted jet centerline velocity was approximately 50% of confirmatory analyses supporting licensing decisions.
the measured value at multiple measurement locations. This Benchmarking activities are one way in which the NRC can
velocity benchmark is likely affected by the difference in the He maintain expertise and proficiency in CFD methods to ensure
content between the experiment and prediction. The predicted that our regulatory activities are effective and efficient.
jet centerline velocity with the one-sigma uncertainty interval One example of the NRC’s efforts to develop and maintain
did not bound the experimental data. expertise in CFD applications comes from our participation in
Keywords: Uncertainty, CFD, Stratified, Mixing the OECD/NEA sponsored PANDA benchmark exercise [2]. In
this blind benchmark, a stratified helium layer is eroded from
1. INTRODUCTION below due to a rising turbulent jet. This benchmark is geared
The U.S. Nuclear Regulatory Commission (NRC), like all towards the development of three-dimensional containment
nuclear safety regulators worldwide, is tasked with regulating its modeling techniques where the mixing of hydrogen exposed to

This material is declared a work of the U.S. Government and is not subject to copyright protection in the United States.
Approved for public release. Distribution is unlimited.
V001T12A002-1
steam jets is a phenomenon of interest. In the experiment, Note that these conclusions are paraphrased from the synthesis
helium was used instead of hydrogen for safety reasons, and air report and do not necessarily represent the state-of-the-art in
was used instead of steam to limit the significance of CFD analysis.
condensation and simplify the CFD benchmark. The rising jet
becomes negatively buoyant when it encounters the helium layer, 2. OVERVIEW OF PANDA EXPERIMENT

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
and this impacts the turbulence phenomena. These phenomena During the experiment, a helium-air mixture (Re ≈ 20,000)
can be challenging for traditional turbulence models used in CFD is injected into a stainless-steel cylindrical vessel from a 75.3
codes and turbulent mixing in stratified layers is significant in a mm inner diameter vertical pipe (Figure 1) [2]. The mixture
variety of nuclear safety applications. NRC was an original flowed upwards until it impacted and mixed with a stratified
participant in this blind benchmark study which concluded in layer of helium (starting at an elevation, y, of 6 m). Buoyancy
2016. The purpose of this follow-on effort is to document the forces kept the relatively heavy jet from penetrating through the
NRC’s efforts to utilize this benchmark study to develop low-density helium rich layer. Gas exited the vessel after passing
expertise in modeling and uncertainty methods as well as best through a funnel near the bottom of the vessel. System pressure
practice modeling guidance. was maintained near atmospheric pressure for the duration of the
The results for 19 participants from the blind benchmark, transient. The helium layer was largely eroded within one hour.
including NRC, are summarized in a synthesis report [2].
Participants in the blind benchmark used a variety of CFD codes
and modeling options. Participants significantly overpredicted
the helium layer erosion rate in some cases and underpredicted
the erosion rate in others. In some instances, different
participants used the same code and turbulence model but
obtained significantly different results. As the reporting only
documents mesh size, CFD code, and turbulence model for each
submission, the basis for the variance between simulated results
is not documented. Based on these observations and the
comparison of calculated results, the authors of the synthesis
report made the following (paraphrased) conclusions:
• The authors don’t know the extent that Best Practice
Guidelines (e.g., grid convergence studies, solver
convergence studies, etc.) were applied. It’s unclear if
more rigorous application of these guidelines would have
reduced the spread in the results.
• The spread in calculated values may be due to variations
between participants’ modeling approaches for the jet inlet
boundary condition, turbulent Schmidt number, molecular
diffusivity, and heat transfer between the gas and vessel Figure 1: Vessel from Test Facility at Paul Scherrer
wall. Institute (PANDA). The left schematic is a cross-section at the
• The spread in calculated values may be due to differences z = 0 m plane at the start of the experiment (+y directed up, +x
between the participants’ CFD codes. directed right).
• It’s unclear what Reynolds-averaged Navier-Stokes
(RANS) turbulence model is best suited for the PANDA At the start of the test, velocities in the vessel were zero and
experiment. temperatures varied from 20 °C to 23 °C. There was an initial
• RANS turbulence models may be more appropriate for helium-air layer established in the upper vessel and the vertical
modeling the PANDA facility than large eddy simulations mass fraction distributions at the start of the test are plotted in
(LES). LES generally performed worse than RANS and Figure 2.
were approximately an order of magnitude more The helium and air mass flow rates in the injection line were
computationally expensive. 0.42 g/s (with reported ± 0.0225 g/s 99% confidence interval)
• There is a large variance in calculated velocity and and 21.53 g/s (σ = 0.23 g/s), respectively. A small amount of
turbulent kinetic energy. Modeling the free jet is a water vapor was also present in the injection mixture. Helium,
challenging component of this benchmark. air, and water vapor mole fractions were 0.134 (σ = 0.0027),
0.862 (σ = 0.0027), and 0.004, respectively. The injected flow is
• There is not an established CFD methodology for
expected to become fully developed (the straight region of the
evaluating erosion of hydrogen layers exposed to steam
pipe before injection was 2.3 m, L/D = 30.5) prior to exiting the
jets, even for the present simplified conditions.
inlet pipe. The temperature of the injection flow was 23°C at the
start of the test, 26°C after 15 seconds, and temperature increased
to 29°C over the remainder of the test.

V001T12A002-2
The jet velocity was measured just above the inlet pipe exit pipe axis was located at x = -650 mm, so these points are directly
using particle imagery velocimetry (PIV). The measurements above the jet exit which was at an elevation of y = 2.995m. The
were made in a separate test with similar mass flow with estimated temperature measurement and position uncertainty
temperature estimated to be 15°C (it wasn’t reported). The PIV were ±0.7 K and ±5 mm, respectively.
data showed that the jet profile 7.3 mm above the inlet pipe exit Helium mole fraction data were obtained from mass

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
plane is close to being fully symmetric along the x-axis. Figure spectrometer measurements at 19 locations. 18 locations were
3 shows the velocity profile above the jet exit (centered on x = - reported at staggered times with an interval between
0.65 m) with the left half of the profile mirrored over to the right measurements of 226 seconds; the 19th location was reported
side to highlight the small deviation from full symmetry. The every 30 seconds. The estimated measurement and position
average mixture exit velocity was approximately 4.67 m/s. The uncertainties for the spectrometer were ±0.5% and ±5 mm,
Froude number was approximately 14. respectively. In addition, the experimenters reported the time at
which the measured helium mole fraction dropped below 0.2 for
10 locations above the jet exit. This data was used to illustrate
the erosion rate of the helium layer and was a key benchmark
8.0 parameter.
PIV velocity and RMS velocity measurements were
reported for seven horizontal lines and three vertical lines near
y-position (m)

6.0
the boundary of the helium rich layer. Measurements were
obtained at t = 111, 715, 1213, 1795, and 2550 seconds.
Turbulent kinetic energy data was also provided along a vertical
4.0
line above the jet exit at t = 111, 1213, and 2550 seconds. The
experimenters did not report on PIV measurement uncertainty.
Air
2.0 He
3. APPROACH AND BASE MODEL
H2O
One purpose for this work is to develop best practice
0.0 guidance for CFD modeling of this type of phenomena by
0.0
0.4 0.2
0.6 0.8 1.0 comparing predictions from CFD models to experimental data.
He Mole Fraction The methodology outlined in the ASME V&V 20-2009 Standard
Figure 2: Initial vessel species concentrations. [3] is used to assess the accuracy of the CFD models and to
estimate the uncertainty for the base CFD model. ANSYS Fluent
6.0 V19.0 was used for all simulations unless specified otherwise.
The initial step was the development of a base model that
served as a reference point for comparison to other models. The
5.0 base model was developed based on experience gained during
the original blind benchmark exercise, where ANSYS Fluent
y-velocity (m/s)

V15.0 was used, and a more recent series of sensitivity studies.


The base model parameters and settings that were established are
4.0
summarized in Table 1 through Table 4. To ensure iterative
convergence, 50 iterations were used for each time step. This
value was increased at the beginning of the simulation to 80
3.0 iterations per time step for the first 0.7 s, and 70 iterations per
time step from 0.7 to 8 s.
Only the vertical straight section, L/D = 30.4, of the inlet
2.0 pipe was modeled. This section is considered long enough to
-0.69 -0.67 -0.65 -0.63 -0.61 establish fully developed flow and the flow upstream of this
x-position (m) section was not considered. The inlet mass flow rate, which was
specified at the bottom of this 2.3 m section of pipe, was
Figure 3: Jet exit velocity. The black dots are values from established by adjusting the inlet velocity specification to match
-0.69 m < x < -0.65 m reflected about the jet axis. the experimental inlet pipe mass flow values at the given
temperature and gas density. The model used the measured
Temperature, mass spectrometry, and velocity data were initial temperature and helium distributions as initial conditions
measured and reported for this benchmark. This data was used for the vessel.
for assessing the CFD predictions. All measurements used were Once the base case was established, sensitivity studies were
on the z = 0 plane. The experimenters reported thermocouple conducted to investigate the model sensitivity to the grid, time
measurements at five locations ([x,y] = [-650mm, 7478mm], [- step, number of iterations within each time step, turbulence
650,4326], [ -650, 3676], [-650, 3036], [-325, 3676]). The inlet model, turbulent Schmidt number, inlet mass flow, model

V001T12A002-3
options (e.g., wall treatment and species diffusion options), and Table 3: Solution methods for base model.
solver options (e.g., spatial gradient discretization
methodology). Based on experience gained during the original Pressure-Velocity Coupling
SIMPLE
blind benchmark and documentation in the associated synthesis Scheme
report, it was assumed the erosion rate and jet velocity were not Green-Gauss Node

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
Spatial Gradient Discretization
sensitive to ambient heat losses; ambient heat losses were not Based
evaluated in the current set of sensitivity studies. Pressure Spatial Discretization Body Force Weighted
The simulation results from the sensitivity studies were Momentum, TKE, TDR, H2O, He,
Second Order Upwind
ranked using the ranking algorithm described in the benchmark & Energy Spatial Discretizations
synthesis report [2]. The algorithm ranks simulations by Transient Formulation Second Order Implicit
summing the absolute difference between calculated and Warped-Face
measured values. Ranking criteria include the helium erosion Other Options Gradient Correction
rate (referred to as “He Drop”), helium concentration time On
histories, jet velocity, local temperature, and a global ranking that
scaled, weighted, and summed the other rankings. The Table 4: Under relaxing factors (URFs) for base model.
simulation rankings were used to help develop best practice
guidelines. Pressure URF 0.3
Simulation uncertainty due to numerical uncertainty and Momentum URF 0.65
input parameter uncertainty was estimated using the grid Density, Body Forces, Turbulent
convergence index (GCI) and a sensitivity coefficient approach, 1.0
Viscosity, H2O, He, & Energy URFs
respectively. In this context, simulation input parameters are TKE & TDR URFs 0.7
continuous parameters input by the user (e.g., turbulent Schmidt
number).
4. GRIDS AND NUMERICAL UNCERTAINTY
Table 1: General model information for base case. Multiple grids were considered to optimize the grid design
for efficiency and accuracy. The final grid for the base case used
Solver ANSYS Fluent V19.0, RANS (Transient) 1.5 million cells (see Figure 4). A grid convergence index (GCI)
Turbulence k-ε Realizable, Enhanced Wall Treatment, [3, 4, 5, 6] method was used to estimate convergence and
Model Full Buoyance Effects numerical uncertainty. GCI was calculated using coarse (0.5M),
Sct 0.7 base (1.5M), and fine (4.5M cells) grids.
Energy
Yes
Equation
Wall Heat
No
Transfer
Species Transport with Diffusion Energy
Species Source (Full Multicomponent Diffusion is
not used)
Gravity Yes
Unstructured hexagonal (inlet pipe) and
Mesh
polyhedral (vessel), 1.5 Million Cells

Table 2: Boundary conditions for base model.

Boundary Boundary Condition


Walls Smooth, No Slip Option, Adiabatic
Uniform Velocity Inlet
Figure 4: Grid on Vertical z=0 Plane.
Turbulent Intensity = 5%
Jet Inlet Hydraulic Diameter = 0.0753 m The coarse grid was produced using the ANSYS meshing
Based on Transient Experimental mass flow platform. The first step was the assignment of grid sizing over
and Temperature several regions of the model. The grid size was reduced in
Outlet Pressure Outlet regions where high gradients are expected and increased in
quiescent regions of the domain (see Figure 4). Tetrahedral cells
were applied for all regions except for the inlet pipe. The inlet
pipe was created using extruded cells from the inlet consisting of
quad elements along the wall and triangular elements in the

V001T12A002-4
central region of the pipe. The final step in the coarse grid numerical solution is φ2 ± GCIBase. Calculated confidence
creation was the conversion of the tetrahedral cells to polyhedral intervals are plotted on the base case results in the right column
using the conversion routine in the FLUENT solver. of Figure 5.
This process was repeated for the base (medium) grid with 0.3
all tetrahedral regions assigned a (refined) grid size reduced by a (a)

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
factor of 1.4. The same procedures were used and all tetrahedral
cells were converted to polyhedral using the FLUENT solver. 0.2

y-velocity (m/s)
The hexagonal grid cells on the inlet pipe wall were not refined
for any of the grids in order maintain consistent y+ values within 0.1
the inlet pipe between models (inlet pipe y+ ≈ 24, targeted at 30,
for all models). The finest grid was developed by refining each 0.0
region of the medium grid by the same (~1.4) factor.
The left column of Figure 5 illustrates the predicted solution -0.1
from the three separate grids and helps to demonstrate the 0.2 0.4 0.6 0.8 1.0 1.2 0.4 0.6 0.8 1.0 1.2
convergence of the solution as the grid is refined. The plots in x-position (m)
the right column of Figure 5 show the results of the base case
with GCI included. It is noted that the predictions were (b)
6.2
considered to be converged with respect to the time step which

y-position (m)
was held constant for each case. See the time step convergence
study notes below. 6.0
GCI is based on Richardson Extrapolation where it is
assumed that the calculated solution will asymptotically Fine
approach the converged numerical solution with successive grid 5.8
Base
refinements. The calculated values for successive grid Coarse
refinements may be extrapolated to predict the fully converged 5.6
numerical solution with the following equation. 0.1 0.2 0.3 0.4 0.5 0.1 0.2 0.3 0.4 0.5
y-velocity (m/s)
𝜑𝜑𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 = 𝜑𝜑1 +
𝜑𝜑1 −𝜑𝜑2
(1) 2500
𝑟𝑟 𝑝𝑝 −1 Fine (c)
2000 Base
Where φ1 and φ2 are the solutions for the fine and base grid, Coarse
respectively. r is the grid refinement factor. p is the order of 1500
Time (s)

convergence, which is calculated from the results using (2).


1000
𝑝𝑝 = [1⁄ln(𝑟𝑟)][𝑙𝑙𝑙𝑙|𝜀𝜀32 /𝜀𝜀21 |] (2)
500
Where ε32 = φ3 – φ2. Using this theory and the guidance
0
outlined in ASME V&V 20 [3], the GCI was calculated for the 6.0 6.2 6.4 6.6 6.8 6.2 6.4 6.6 6.8
base (medium) grid using (3). y-position (m)
𝐹𝐹𝐹𝐹∗|𝜀𝜀21 |
𝐺𝐺𝐺𝐺𝐺𝐺𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 = 𝑟𝑟 𝑝𝑝 (3) Figure 5: The left plots are calculated values with grid
(𝑟𝑟 𝑝𝑝 −1)
refinement. The right plots are the base case results with 95%
confidence interval for numerical uncertainty. (a) y-velocity
Where Fs is a safety factor which is set equal to 1.25. The
across jet (y = 5.9 m and t = 715 s). (b) y-velocity along jet
1.25 value was used instead of the more common value of 3.0
axis (x = -0.65 m and t = 1213 s). (c) Time helium mole
because a careful study was completed using 3 grids and the
fraction becomes less than 0.2 (above jet exit, x = -0.65).
value of 3.0 is considered too conservative in this case based on
the authors judgement of the results. A value of 3, however, A time step convergence study was conducted for the base
could easily be justified. Chapter 5 of the book “Fundamentals case. Calculated values were very similar for dt = 0.1 s (base
of Verification and Validation” [5] has a good discussion on the case) and dt = 0.2 s, and numerical uncertainty due to the
selection of the safety factor. The calculated order of timestep is considered to be negligible compared to numerical
convergence, p, varied from approximately 1 to 5 and was on uncertainty due to the spatial grid. For example, the average
average around 2. p = 2 was used for all points in Equation 3. difference between calculated values for the fine and base grid
The values of φ were interpolated to the measurement locations in Figure 5 (c) is 60 s; the average difference for dt = 0.1 s and
using a user defined function in the FLUENT solver which used dt = 0.2 s is 0.4 s. In addition, the numerical uncertainty due to
the solution gradients to obtain an approximately second order convergence within each time step was assumed to be negligible.
interpolation. The 95% confidence interval for the exact

V001T12A002-5
The solutions were very similar when the number of iterations experimental data, but the gradient in helium concentration was
per time step was reduced by 10 (see Table 5 and Table 6). similar to the experimental data. Finally, the k-ε RNG turbulence
model results were similar to the base case and experimental
5. SENSITIVITY STUDIES data, except the helium layer diffused more slowly.
Predictions were completed to investigate the model

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
sensitivity to the turbulence model, turbulent Schmidt number
(Sct), inlet mass flow rate, model options (e.g., wall treatment
and species diffusion options), and solution options (e.g., spatial
gradient discretization methodology). These simulations are
summarized in Table 5.
The sensitivity studies were ranked using the ranking
algorithm described in [2]. Rankings are presented in Table 6.
The helium mole fraction at x = 0.3 m and t = 111, 1213, and
2550 s is plotted for different turbulence models in Figure 6. The t = 111 s
helium mole fraction data was measured at staggered time
intervals of 226 s. Data for Figure 6 was obtained by linearly
interpolating between the two closest staggered measurements,
and the uncertainty bars are the difference between these two
measured values. Data in Figure 6 were at multiple x-locations.
Full contours of the helium mole fraction on a vertical plane are
shown in Figure 7.
8 t = 111 s t = 1213 s t = 2550 s
t = 1213 s

6
y-position (m)

k-ε Rlz
k-ω Std
k-ω BSL t = 2550 s
2
k-ω SAS
k-ω SST k-ε Realizable k-ω-SST k-ω Standard
k-ε RNG
Data 0.37 0.185 0.0
0
0.0 0.1 0.2 0.3
0.1 0.2 0.3 0.1 0.2 0.3 0.4
Figure 7: Helium model fraction contours for k-ε Realizable
Helium Mole Fraction
(base case), k-ω-SST (R12), and k-ω-Standard turbulence
Figure 6: Helium mole fraction for various turbulence models.
models.
The y-velocity and y-RMS velocity for y = 5.1 m and t = 111
The base case (R03, k-ε Rlz) had the best mass spectrometry s are plotted in Figure 8. This data was selected for comparison
(MS, i.e., helium) ranking and appeared to have the best because it is early in the transient and located in the relatively
agreement with the experimental data in Figure 6. k-ω BSL air-rich region two meters above the jet exit (see Figure 6).
(R23) and k-ω SST (R12) turbulence models resulted in faster However, the calculated helium mole fraction varies by
diffusion and a smoother gradient in the helium concentration approximately 0.05 between the turbulence models in this
than observed in the experiment. The k-ω Standard (R22) region, which will likely affect the velocity. The base case (R03,
turbulence model was similar to the base case at t = 111s, but k-ε Realizable) under predicted the peak velocity. The velocity
then rapidly diffused and resembled k-ω BSL (R23) and k-ω SST and RMS velocity for the k-ω SAS (R23) and k-ω standard (R22)
(R12) turbulence models at later times. This is likely due to the turbulence models were a significantly different shape than the
first order differencing used for the k-ω Standard (R22) other turbulence models. The sharp peak in the k-ω SAS
simulation (2nd order differencing schemes were unstable). k-ω turbulence model likely contributed to the faster diffusion rate
SAS (R10) turbulence model resulted in faster diffusion than the

V001T12A002-6
observed in Figure 6. The k-ω standard peak velocity was most faster helium diffusion, less peaked velocity above the jet for all
similar to the experimental data, but again these results are measurement times.
suspect due to the use of the first order differencing scheme. While participating in the blind benchmark, it was observed
that using second order pressure discretization in FLUENT
2.0 V15.0 resulted in unphysical turbulent viscosity ratios

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
(a) throughout the domain and rapid dissipation of the helium layer.
k-ε Rlz
k-ω Std Sensitivity studies were conducted in the present study using
1.5 FLUENT V19.0 to reevaluate second order pressure
k-ω BSL
discretization. The results using second order spatial
y-Velocity (m/s)

k-ω SAS
1.0 k-ω SST discretization (R27) were very similar to the base case (body
k-ε RNG force weighted) using FLUENT V19.0. It is assumed that this
Data previous poor result from the second order pressure
0.5 discretization method has been corrected in the FLUENT solver.
It is noted, however, that a hexagonal grid was used for the earlier
blind benchmark and a polyhedral grid is used for the present
0.0
study.
Rapid helium diffusion was also observed using cell based
(b) spatial gradient discretizations in FLUENT V15.0. Cell based
0.3
y-RMS velocity (m/s)

methods were reevaluated for FLUENT V19.0 in cases R24


(Green-Gauss cell based) and R25 (Least Squares cell based).
The results for node based (R03), Green-Gauss cell based, and
0.2 Least Squares cell based were very similar in the present study.
Both the cell based methodologies ranked slightly better than the
base case. The Least Squares approach predicted slightly faster
0.1 helium diffusion than the other two approaches, and slightly
better agreement with the data.
The coupled scheme for pressure-velocity coupling resulted
in unphysical helium diffusion. No further testing with this
0.0 method was attempted and it is not recommended at this time for
-1.2 -1.0 -0.8 -0.6 -0.4 -0.2 the PANDA experiment.
x-position (m) In addition to these studies on modeling options, studies
were also conducted to evaluate solution sensitivity to boundary
Figure 8: (a) y-velocity and (b) y-RMS velocity at y = 5.1
conditions and other physical parameters.
m and t = 111 s.
Uncertainty in the inlet mass flow rate caused uncertainty in
y+ ≈ 24 (on average) was observed for the inlet pipe wall for the final calculated solution. The 99% confidence interval for
the base (R03), coarse (R02), and fine grids (R04). Recall that the helium mass flow rate was 0.42 ±0.0225 g/s. The confidence
all grids had the same wall cell size on the inlet pipe wall. The interval for the air mass flow rate was not provided, but the
y+ values for the inlet pipe wall treatment sensitivity studies standard deviation in measured flow was 0.23 g/s (the nominal
(R08 and R09) were also approximately 24. On the vessel outlet air flow rate was 21.53 g/s) [2]. As a result, it is hard to quantify
pipe, y+ values were approximately 50. Flow near the vessel the total uncertainty in the inlet mass flow rate.
outer wall boundaries was generally quiescent and the median The inlet mass flow rate for the base case was adjusted by
y+ values over the outer walls was less than 5. This outer wall ±5% to evaluate solution sensitivity to variations in the inlet
region is not considered significant for the jet mixing and mass flow rate. Calculated results and PIV measurements for jet
interactions with the helium layer which occur in central regions exit velocity (7.3 mm above the jet exit) are plotted in Figure 9.
of the vessel. The measured peak velocity at the jet exit is 5.35 m/s compared
The differences in helium and velocity predictions between to only 5.0 m/s for the CFD case with 5% increase in the inlet
the base case and wall treatment sensitivity studies (R08 and mass flow rate (R32).
R09) were very small. The simulation using scalable wall Assuming reasonable uncertainty in the measured velocity,
functions (R09) had slightly faster helium diffusion and slightly the ±5% variation in inlet mass flow rate could not account for
more peaked velocity than the base case and the simulation using the difference between the calculated and measured values
the Menter-Lechner wall treatment option. R09 had the best (Figure 10). Predictably, the increased mass flow rate caused
ranking of the wall treatment sensitivity studies by a small the helium layer to erode more rapidly. These results are
margin. plotted in Figure 11.
The base case results agreed with the data better than the full Multiple additional sensitivity studies were conducted
multicomponent diffusion case (R15). R15 had significantly where the inlet temperature, species mole fractions, and pressure
were adjusted. Reasonable variations in these parameters could

V001T12A002-7
not account for the difference in the measured and calculated 8
peak velocities. The experimental velocities suggest that the t = 111 s t = 1213 s t = 2550 s
mass flow is higher than reported. This issue has not been fully
resolved.
6.0

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
6

y-position (m)
5.0
y-velocity (m/s)

4
4.0

R03 (base)
R32 (+5%) 2
3.0 R03 (base)
R33 (-5%) R32 (+5%)
Data R33 (-5%)
Data
2.0 0
-0.69
-0.65 -0.63 -0.61 -0.67 0.0 0.1 0.2 0.3
0.1 0.2 0.3 0.1 0.2 0.3 0.4
x-position (m) Helium Mole Fraction
Figure 9: Measured jet exit velocity compared to Figure 11: Helium mole fraction for varied inlet mass flow
calculated jet exit velocity with +5%, -5%, and nominal inlet rates.
mass flow.
The turbulent Schmidt number (Sct) is the ratio of the
1.0 momentum and mass diffusivity. The default value in the
R03 (base) (a)
R32 (+5%) ANSYS k-ε Realizable turbulence model formulation used is
0.8
R33 (-5%) 0.7. This value, however, is known to vary significantly [7].
Data Estimates of this variation were made since values for this
0.6
y-Velocity (m/s)

particular flow field are unknown. An assumed uncertainty in


the Sct contributed to uncertainty in the calculated helium
0.4
diffusion rate, as shown in Figure 12. The variation in y-velocity
and y-RMS velocity when varying Sct from 0.35 to 1.4 was
0.2
similar to the spread observed in Figure 10 (although less linear).
0.0 8 t = 111 s t = 1213 s t = 2550 s

(b)
0.3 6
y-RMS velocity (m/s)

y-position (m)

0.2
4

0.1
Sct=0.7 (R03)
2 Sct=1.4 (R06)
0.0 Sct=1.05 (R07)
-1.2 -1.0 -0.8 -0.6 -0.4 -0.2 Sct=0.35 (R31)
x-position (m) Data
0
Figure 10: (a) y-velocity and (b) y-RMS velocity at y = 5.1 0.0 0.1 0.2 0.3
0.1 0.2 0.3 0.1 0.2 0.3 0.4
m and t = 111 s. Helium Mole Fraction
Figure 12: Helium mole fraction for varied turbulent
Schmidt number.

V001T12A002-8
Table 5: Key for sensitivity studies. Runs generally used Table 6: Ranking results of sensitivity studies. The ranking
the model defaults and the values specified for the base model is a measure of the difference between the measured and
unless specified below. predicted value. Smaller values imply better agreement with
data.
Run

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
Sensitivity
ID Run Ranking
k-ε Realizable Turbulence Model Sensitivities ID Global He Drop PIV MS T
R02 Coarse mesh (0.5 M cells) [2]* 0.0324 132.13 0.031 0.0064 0.33
R03 Base case (see Table 2 through Table 4) k-ε Realizable Turbulence Model Sensitivities
R04 Fine mesh (4.5 M cells) R02 0.0645 132.56 0.158 0.0062 0.72
R05 dt = 0.2 s R03 0.0607 85.75 0.153 0.0063 0.71
R06 Sct = 1.4 R04 0.0600 78.54 0.152 0.0066 0.71
R07 Sct = 1.05 R05 0.0608 85.76 0.154 0.0063 0.71
R31 Sct = 0.35 R06 0.0836 313.47 0.165 0.0122 0.73
R32 Increase inlet mass flow by 5% R07 0.0717 189.09 0.159 0.0094 0.72
R33 Decrease inlet mass flow by 5% R31 0.0630 167.94 0.151 0.0051 0.68
R08 Wall treatment: Menter-Lechner R32 0.0556 94.06 0.138 0.0044 0.72
R09 Wall treatment: Scalable Wall Functions R33 0.0793 245.53 0.172 0.0109 0.69
R15 Use Full Multicomponent Diffusion R08 0.0621 88.88 0.156 0.0064 0.74
R16 Warped-Face Gradient is off R09 0.0592 76.33 0.149 0.0063 0.72
Spatial Gradient Discretization: Green-Gauss cell R15 0.0821 254.16 0.186 0.0097 0.70
R24
based R16 0.0607 85.45 0.154 0.0063 0.71
Spatial Gradient Discretization: Least Squares cell R24 0.0594 73.00 0.153 0.0060 0.71
R25
based R25 0.0590 68.76 0.152 0.0060 0.71
R27 Pressure Spatial Discretization: Second Order R27 0.0601 79.32 0.152 0.0063 0.72
R28 Pressure-Velocity Coupling: Coupled Scheme R28 Poor results. Unstable solution
R34 Number of iterations per time step reduced by 10 R34 0.0607 85.46 0.154 0.0063 0.71
Turbulence Model Sensitivities Turbulence Model Sensitivities
k-ω Standard turbulence with k-ω wall treatment R22 0.0680 138.54 0.149 0.0122 0.64
-1st order upwind spatial discretizations* R26 0.0752 125.79 0.177 0.0136 0.61
R22
-Production limiter, shear flow and low-Re R23 0.1079 437.21 0.204 0.0221 0.51
corrections R10 0.1181 525.84 0.227 0.0129 1.10
R26 R22 except Low-Re corrections are off R12 0.1073 437.06 0.202 0.0219 0.52
k-ω-BSL turbulence with k-ω wall treatment R14 0.0800 232.69 0.183 0.0093 0.71
R23
-Low-Re correction and production limiter R15 0.0821 254.16 0.186 0.0097 0.70
k-ω Standard with SAS & k-ω wall treatment *These results are the highest ranked results per category from the
R10 -Least squares cell based & bounded 2nd order imp. synthesis report.
-Production limiter and shear flow corrections
k-ω-SST turbulence with k-ω wall treatment 6. UNCERTAINTY QUANTIFICATION
R12
-Production limiter and shear flow corrections Most of the sensitivity studies completed were focused on
k-ε-RNG modeling and solver options or the various turbulence models
R14 -PRESTO! pressure discretization available. This approach helped to demonstrate the applicability
-Differential viscous model of the base case approach. These sensitivity studies, however,
R15 R14 except differential viscous model is off do not directly feed into the uncertainty of the base case. Two
*Second order differencing schemes were unstable for k-ω parameters that clearly impact the base case model are the mass
Standard. flow and turbulent Schmidt number. Solution uncertainty for
the base case (R03) due to prescribed uncertainty in these input
parameters was calculated using the sensitivity coefficient
method outlined in the ASME V&V 20-2009 standard [3]. This
input uncertainty is calculated using (4).

𝜕𝜕𝜕𝜕 2
2
𝑢𝑢𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 = ∑𝑛𝑛𝑖𝑖=1 � 𝑢𝑢𝑋𝑋𝑖𝑖 � (4)
𝜕𝜕𝑋𝑋𝑖𝑖

V001T12A002-9
Where S is the local simulation result and 𝑢𝑢𝑋𝑋𝑖𝑖 is the standard
uncertainty in input parameter Xi. Uncertainty was estimated 8
(a) (b) R03 (base)
using the two input parameters (i.e., n = 2) as a demonstration of
the method. More parameters could be considered with the same Data
approach if the input parameter and its associated uncertainty

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
(c)
were identified. The first input parameter is the inlet mass flow 6
rate with an estimated standard uncertainty of ±5% of the

y-position (m)
reported value. The second input parameter is the turbulent
Schmidt number (Sct). The standard uncertainty for Sct is
estimated to be ±0.35 (around the default value of 0.7). The 4
derivative in (4) is calculated using the second-order accurate 0.2 0.5 0.8 0.175 0.25 0.325
y-velocity y-RMS velocity
finite difference formulation defined in (5) and CFD predictions
(m/s) (m/s)
R07, R31, R32, and R33 (see Table 5).
1.0

y-velocity (m/s)
2 (d)
𝜕𝜕𝜕𝜕
=
𝜕𝜕(𝑋𝑋𝑖𝑖 +∆𝑋𝑋𝑖𝑖 )−𝜕𝜕(𝑋𝑋𝑖𝑖 −∆𝑋𝑋𝑖𝑖 )
(5) 0.6
𝜕𝜕𝑋𝑋𝑖𝑖 2∆𝑋𝑋𝑖𝑖
R03 (base) 0.2
The sensitivity coefficient method was selected because it is Data
0 -0.2
significantly less computationally expensive than sampling-
0.0 0.1 0.2 0.3 0.4 -1.3 -1.0 -0.7 -0.4 -0.1
based methods such as Monte Carlo techniques as noted in other
studies [3, 6]. It is noted, however, that the sensitivity coefficient Helium Mole Fraction x-position (m)
approach is most applicable when sensitivities show a linear
variation. Figure 13: Experimental measurements and calculated
The input parameter uncertainty and numerical uncertainty results for t = 111 s. (a) Calculated helium mole fraction at x =
are combined using (6). un,i is the one-sigma (68%) solution 0.3 m. (b) y-velocity above the jet exit (x = -0.65 m). (c) y-
uncertainty due to numerical and input parameter uncertainty; it RMS velocity above jet exit. (d) y-velocity at y = 5.1 m.
does not include uncertainty due to model assumptions such as
the assumptions built into the k-ε Realizable turbulence model.
In accordance with the ASME V&V 20-2009 standard, the GCI
8
95% confidence intervals are converted to the one-sigma (a) (b) R03 (base)
uncertainty interval by dividing GCIbase by 1.15. Data
2 2
𝑢𝑢𝑖𝑖,𝑖𝑖 2
= 𝑢𝑢𝑖𝑖𝑖𝑖𝑛𝑛 + 𝑢𝑢𝑖𝑖𝑖𝑖𝑒𝑒𝑖𝑖𝑒𝑒 (6) (c)
6
The base case calculated results and one-sigma uncertainty
y-position (m)

interval (S ± un,i) for t = 111, 1213, and 2550 s are plotted with
the experimental data in Figure 13, Figure 14, and Figure 15,
4
respectively. The calculated helium mole fraction agrees with 0.2 0.5 0.8 0.175 0.25 0.325
the data well. The available data is generally bounded by the y-velocity y-RMS velocity
calculated results and the one-sigma uncertainty interval. The (m/s) (m/s)
erosion of the He layer is well predicted. Uncertainties are 1.0
y-velocity (m/s)

highest in the region where the He layer mixing front is located 2 (d)
0.6
which is the region of highest He concentration gradients.
R03 (base) 0.2
Data
0 -0.2
0.0 0.1 0.2 0.3 0.4 -1.3 -1.0 -0.7 -0.4 -0.1
Helium Mole Fraction x-position (m)
Figure 14: Experimental measurements and calculated
results for t = 1213 s. (a) Calculated helium mole fraction at x
= 0.3 m. (b) y-velocity above the jet exit (x = -0.65 m). (c) y-
RMS velocity above jet exit. (d) y-velocity at y = 5.9 m. The
local decrease in experimental velocity near x = -0.7 m is
probably due to unidentified material in the domain.

V001T12A002-10
and measured values are due to some combination of modeling
8 error (e.g., assumptions in the k-ε Realizable turbulence model),
(a) (b) (c) measurement error (velocity measurement uncertainty was not
provided, so this can’t be quantified), and input and numerical
error greater than the one-sigma confidence interval.

Downloaded from http://asmedigitalcollection.asme.org/VVS/proceedings-pdf/VVS2020/83594/V001T12A002/6554388/v001t12a002-vvs2020-8821.pdf by Luis Jean Pierre Jacobo Fuentes on 21 October 2020
6
7. CONCLUSION
R03 (base)
y-position (m)

CFD calculations were conducted and compared to


Data
experimental data of erosion of a stratified helium layer due to
Data
4 the influence of a weakly turbulent jet to develop best practice
0.2 0.5 0.8 0.075 0.15 0.225 guidelines and to estimate the uncertainty in the CFD
y-velocity y-RMS velocity predictions.
(m/s) (m/s) The k-ε Realizable turbulence model with second order
1.0
y-velocity (m/s)

2 differencing is proposed as the method which results in the best


0.6 agreement with the data. The enhanced wall treatment was used
with good results.
0.2 The numerical error is estimated, and this provides
-0.2 confidence in the convergence of the solution. CFD uncertainty
0
0.0 0.1 0.2 0.3 0.4 -1.3 -1.0 -0.7 -0.4 -0.1 due to the numerical error and key input uncertainties was
calculated using a Grid Convergence Index (GCI) method and
Helium Mole Fraction x-position (m) sensitivity coefficient approach, respectively. A total uncertainty
was computed from the combination of both components. The
Figure 15: Experimental measurements and calculated
predictions of helium concentration erosion showed good
results for t = 2550 s. (a) Calculated helium mole fraction at x
agreement with the data which was generally within by the CFD
= 0.3 m. (b) y-velocity above the jet exit (x = -0.65 m). (c) y-
one-sigma uncertainty interval. The predicted velocity
RMS velocity above jet exit. (d) y-velocity at y = 6.45 m.
measurements were inconsistent with the test data and the
The peak velocity is generally underpredicted. Even with reasons behind this variation have not been fully identified. The
the uncertainty bound added to the peak velocity, the data are still impact of small variations in the He profiles as well as potential
higher than the predicted range. For example, at t = 111 s, the variations in the mass flow itself are potential sources of this
calculated jet velocity is on average 0.33 m/s less than the error. The data are too sparse to fully quantify this issue.
measured value, and the one-sigma uncertainty interval is
approximately 0.09 m/s (see Figure 13 (b)). The calculated peak REFERENCES
velocity is approximately 50% of the measured value at multiple [1] Nuclear Energy Agency. The Characteristics of an
locations. Effective Nuclear Regulator, NEA/CNRA/R(2014)3, 2014
The difference between the calculated and measured [2] Nuclear Energy Agency. The Nuclear Energy Agency—
velocity is greatest at t = 111 s. Here, the velocity was measured Paul Scherrer Institut Computation Fluid Dynamics Benchmark
just as the jet entered the helium rich region (see Figure 13 (a) Exercise, 2016
and (b)). The jet’s velocity decreases rapidly in this region due [3] The American Society of Mechanical Engineers
to the negative buoyancy forces (i.e., the jet is denser than the (ASME). Standard for Verification and Validation in
helium rich layer), and the jet velocity is therefore sensitive to Computational Fluid Dynamics and Heat Transfer (ASME V&V
the local helium concentration. As helium mole fraction data is 20-2009), 2009
not available near the velocity measurements, it’s unclear if the [4] Roache, Patrick. Perspective: A Method for Uniform
difference between the measured and calculated velocity can be Reporting of Grid Refinement Studies, 1994
attributed to differences in the local helium concentration. [5] Roache, Patrick. Fundamentals of Verification and
The calculated peak velocity and uncertainty interval at t = Validation, Hermosa, 2009
1213 s and t = 2550 s were closer to the experimental data than [6] Boyd, C, et. al. . CFD Modeling Uncertainty for Mixing
at t = 111 s. As seen in Figure 14 (a) and Figure 15 (a), the Benchmark, REMOO Energy, 8th International Conference and
calculated helium mole fraction is slightly overpredicted near the Workshop, May 2018.
jet measurements at these times and it follows that calculated [7] Gualtieri C, et. al. . On the Values for the Turbulent
velocity is less than the experimental value. Schmidt Number in Environmental Flows, Fluids 2017, 2, 17,
The results where the data is not bounded by the calculated 2017.
value ± un,i indicate that the one-sigma (68%) numerical and
input uncertainties can’t fully account for the differences
between the calculated and measured peak velocity. As a result,
it can be concluded that the differences between the calculated

V001T12A002-11

You might also like