You are on page 1of 10

Learning an Eddy Viscosity

Jaideep Ray1 Model Using Shrinkage


Sandia National Laboratories,
MS 9159, P.O. Box 969,
Livermore, CA 94550
and Bayesian Calibration:
e-mail: jairay@sandia.gov
A Jet-in-Crossflow Case Study
Sophia Lefantzi
Sandia National Laboratories, We demonstrate a statistical procedure for learning a high-order eddy viscosity model
MS 9152, P.O. Box 969, (EVM) from experimental data and using it to improve the predictive skill of a Reynolds-
Livermore, CA 94550 averaged Navier–Stokes (RANS) simulator. The method is tested in a three-dimensional
e-mail: slefant@sandia.gov (3D), transonic jet-in-crossflow (JIC) configuration. The process starts with a cubic eddy
viscosity model (CEVM) developed for incompressible flows. It is fitted to limited experi-
Srinivasan Arunajatesan mental JIC data using shrinkage regression. The shrinkage process removes all the terms
Sandia National Laboratories, from the model, except an intercept, a linear term, and a quadratic one involving the
MS 0825, P.O. Box 5800, square of the vorticity. The shrunk eddy viscosity model is implemented in an RANS simu-
Albuquerque, NM 87185 lator and calibrated, using vorticity measurements, to infer three parameters. The cali-
e-mail: sarunaj@sandia.gov bration is Bayesian and is solved using a Markov chain Monte Carlo (MCMC) method. A
3D probability density distribution for the inferred parameters is constructed, thus quan-
Lawrence Dechant tifying the uncertainty in the estimate. The phenomenal cost of using a 3D flow simulator
Sandia National Laboratories, inside an MCMC loop is mitigated by using surrogate models (“curve-fits”). A support
MS 0825, P.O. Box 5800, vector machine classifier (SVMC) is used to impose our prior belief regarding parameter
Albuquerque, NM 87185 values, specifically to exclude nonphysical parameter combinations. The calibrated
e-mail: ljdecha@sandia.gov model is compared, in terms of its predictive skill, to simulations using uncalibrated lin-
ear and CEVMs. We find that the calibrated model, with one quadratic term, is more
accurate than the uncalibrated simulator. The model is also checked at a flow condition
at which the model was not calibrated. [DOI: 10.1115/1.4037557]

1 Introduction higher-order terms, e.g., quadratic and cubic terms [5], but they in
turn introduce more parameters that require calibration. With lim-
Reynolds-averaged Navier–Stokes (RANS) models are the
ited experimental data, properly calibrating all parameters
workhorse of aerodynamic design calculations due to their com-
becomes impossible because of the danger of overfitting.
putational speed vis-a-vis other simulation methods, e.g., large
In this work, we devise a principled procedure for enriching the
eddy simulations (LES). RANS obtains its computational celerity
LEVM and calibrating (tuning) parameters to obtain a predictive
by approximating many of the turbulent processes in the flow.
RANS model for a high Reynolds number interaction of a com-
The approximations include equations governing the evolution of
pressible (Mach number, M ¼ 3.73) jet with a transonic (M ¼ 0.8)
the turbulent kinetic energy k, its dissipation rate e, and a host of
crossflow. The experiment has been described in Refs. [6–8], and
empirical closure models. The empirical closures also contain
it is poorly simulated by RANS (with LEVM) [9]. We hypothe-
parameters whose values are obtained by calibrating to simple
size that simultaneously enriching the LEVM with higher-order
canonical flows [1,2]; we will refer to these values as the
terms and calibrating the parameters to experimental JIC measure-
“nominal” parameter values. Despite their widespread use, RANS
ments could yield a high-order RANS simulator with predictive
models are not particularly predictive for complex turbulent inter-
skill superior to RANS with LEVM or the cubic eddy viscosity
actions such as jet-in-crossflow (JIC), especially at high Reynolds
model (CEVM) described in Ref. [5]. The technical difficulty is
and Mach numbers. Predictive inaccuracy arises from two causes:
twofold. First, it is unlikely that the experimental measurements
First, suboptimal values of model parameters, simply picked from
are sufficiently informative to allow the estimation of all the coef-
literature, are unlikely to be satisfactory for complex flows. This
ficients in a high-order RANS model (nine parameters for a cubic
can be rectified by tuning parameters to relevant experimental/
model). Consequently, we will devise a rigorous method for
high-fidelity data [3,4]. The second cause, model-form error,
selecting the higher-order terms with which to enrich the LEVM.
arises from “missing physics” such as assumptions of isotropy,
Second, the approximations in the RANS models and the limited
neglect of history effects, simplification of various terms in the
experimental measurements may not allow the estimation of
model equations, and assumptions of linear relationship between
parameters with much certainty. To address this, we will adopt a
the turbulent stresses and velocity field strain rates. Here, we focus
Bayesian calibration approach, as developed by us for calibrating
on the last approximation, namely, the linear relationship between
RANS models [4], and use it to estimate RANS and eddy viscos-
the turbulent stresses and the velocity strain rate. In flows with
ity model (EVM) parameters as a joint probability density func-
strong curvature and anisotropy, this simple relationship (called
tion (JPDF). The JPDF will capture the uncertainty in the
the linear eddy viscosity model (LEVM)) is unlikely to suffice.
estimate. To the best of our knowledge, such a data-driven proce-
The problem can be alleviated by enriching the LEVM with
dure to “discover” and configure a high-order RANS model, com-
mensurate with the information content of an experimental
1 dataset, currently does not exist.
Corresponding author.
Manuscript received July 7, 2016; final manuscript received August 2, 2017;
published online September 7, 2017. Assoc. Editor: Yan Wang. 2 Background
The United States Government retains, and by accepting the article for
publication, the publisher acknowledges that the United States Government retains, a
nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the
2.1 Jet-in-Crossflow. Jets perpendicular to the long axis of
published form of this work, or allow others to do so, for United States Government slender aerodynamic bodies are used to spin the vehicle for vari-
purposes. ous applications. These jets convect downstream and generate a

ASCE-ASME Journal of Risk and Uncertainty in Engineering Systems, MARCH 2018, Vol. 4 / 011001-1
Part B: Mechanical Engineering C 2018 by ASME
Copyright V

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


complex vortex system that can interact with the fins of the vehi- were used as proxies for the three-dimensional (3D) RANS model.
cle. This interaction generates a pressure distribution on the fins, They had to excise a part of the parameter space since it was non-
which can result in the production of torque that can counter the physical. A similar approach was used in our previous work [4]
effects of the jets, thereby reducing their effectiveness. Thus, it is on the estimation of three RANS model parameters and a crude
critical to accurately predict this interaction for effective aerody- estimate of the model-form error for the JIC dataset used in this
namic design of such vehicles. In an engineering design setting, study. This paper describes in detail how we isolated the realistic
these predictions are obtained using RANS models with LEVMs part of the parameter space (called R) and modeled it using a sup-
(for computational tractability). Their predictive accuracy for JIC port vector machine classifier (SVMC). It served as a prior distri-
simulations leaves much to be desired. A previous investigation bution for the parameters being estimated. The 3D RANS model
[9] with k  x RANS models showed that the simulations over- was replaced with polynomial surrogates, with stringent accuracy
predicted turbulent intensities as well as the penetration of the jet requirements that could not always be satisfied; consequently, all
into the crossflow. Further, the predicted vortex system is “fatter,” the locations (“probes”) with measurements could not be used in
indicating that the turbulent diffusion is too large. LES simula- the parameter estimation effort. The surrogates and SVMC were
tions of the same JIC configurations obtained much better agree- combined within an adaptive MCMC algorithm (called delayed
ment with experiments [10,11], but they are too computationally rejection adaptive metropolis, see Ref. [27]) to yield JPDFs. They
expensive for routine design investigations. were tested for predictive accuracy in three different JIC interac-
tions. The same calibration framework and software will be
reused in this study. Note that all the studies mentioned earlier
2.2 Reynolds-averaged Navier–Stokes With Linear Eddy used the LEVM in their RANS simulators and none attempted to
Viscosity Model. We have previously investigated if the predictive enrich it in a data-driven manner; however, the studies in Refs.
skill of RANS simulations of JIC could be improved via calibration [17–24] and [28] performed something similar—they enriched the
[4]. There, as well as in this paper, we use a compressibility- turbulent transport model instead.
corrected version of the k  e RANS model; descriptions with an
embedded LEVM are in Refs. [12,13]. As part of this study, we will
enrich this particular form of the RANS equations with the CEVM 2.3 Nonlinear Eddy Viscosity Models. While previous stud-
in Ref. [5]. Further, we will calibrate two parameters, Ce2 ; Ce1 , in the ies have tried to improve parameter values, they have not
e evolution equation, see Eq. (2) in Ref. [4] for details. The nominal addressed some fundamental shortcomings of the LEVM itself.
values of the parameters in the k  e RANS equations are obtained LEVMs are incapable of accommodating anisotropy in turbulent
from simple flows and there are inconsistencies between their values stresses and are also insensitive to curvature in the mean-flow
in various incarnations of the k  e model. Further, these parameters streamlines, e.g., bending of the jet in JIC and swirl, e.g., counter-
do not agree with values derived from simple incompressible-flow rotating vortex pair (CVP). Quadratic terms (leading to quadratic
experiments either, see Refs. [4,14] for a discussion. This uncertainty eddy viscosity models (QEVMs)) allow anisotropy in the normal
arises from the tendency to treat these parameters as “universal” con- turbulent stresses, while cubic terms sensitize them to swirl and
stants that could be estimated from any flow, whereas in reality, they streamline curvatures. Reviews of QEVM and CEVM (collec-
are tunable constants with flow-dependent optimal values. tively, nonlinear eddy viscosity models (NLEVMs)) as well as
There have been previous efforts to tune RANS parameters intercomparisons within k  e models are available for bluff
using a Bayesian approach (i.e., in the form of a probability den- bodies [29], shock-boundary layer interactions [30], and turbulent
sity function (PDF)) as well as to estimate the model-form error. jets [31]. NLEVMs model the turbulent stress sij as (Eq. (2) in
These were primarily performed for simple flows (flat-plate Ref. [5])
boundary layers, channel flows, etc.) [14,15] due to the enormous
cost of Markov chain Monte Carlo (MCMC; see Ref. [16]) solu- 2 kX 3
tions of the Bayesian inverse problems setup to estimate the sij ¼ u0i u0j ¼ dij k  T Sij þ T cl fl ðS; XÞ
3|fflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflffl} ee l¼1
parameters as PDFs. Investigations into more complicated flows |fflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflffl}
intercept linear
have been performed by assuming that the inaccuracies in RANS quadratic
 2 7 (1)
predictions were entirely due to model-form errors. The studies in k X
Refs. [17–21] augment their turbulence transport model (e.g., the þ T cl fl ðS; XÞ
ee l¼4
evolution equation for turbulent viscosity in the Spalart–Allmaras |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
model, or the equation for k in k  e ones) with a spatially depend- cubic
ent source term that is estimated from LES data via a gradient
descent algorithm. The source terms are thereafter modeled using where T ¼ Cl k2 =ee
the local flow variables using machine-learning methods. The    
same philosophy is employed in Ref. [22] where turbulent stress @ @ @ @
Sij ¼ Ui þ Uj ; Xij ¼ Ui  Uj
anisotropy was modeled using spatial basis functions and cali- @xj @xi @xj @xi
brated using iterated ensemble Kalman filters. Wang et al. [23] rffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffi
developed models of turbulent stress anisotropy in terms of local Se ¼
k Sij Sij
; X e ¼ k Xij Xij (2)
flow variables using random forest models, whereas Wu et al. [24] ee 2 ee 2
!2
developed distance metrics to quantify the difference between a 1=2  
@k eXe
set of training flows and a flow where the trained turbulent models ee ¼ e  2 and Cl ¼ Cl S;
would be used. This distance is used to judge the suitability of the @xj
trained models in the particular use case. Weatheritt and Sandberg
[25] developed a genetic algorithm to generate functional forms Various QEVMs differ in terms of the coefficients c1 ; …; c3 ; also,
that could be used to relate local flow properties to turbulent stress Cl is a constant in most of them. CEVMs too vary in the values
anisotropy. used for cl as well as the functional form for Cl in terms of S and
The use of Bayesian estimation of parameters in a complex X. The coefficients and functional forms are summarized in Refs.
flow, i.e., one that requires the RANS simulator to be replaced [29,30]. In bluff-body flows, NLEVMs produce results closer to
with a surrogate model to overcome the phenomenal cost of experimental measurements, primarily because they simulate the
MCMC solutions, was performed in Ref. [26]. Four k  e parame- stagnation points (irrotational regions of large strain rates) accu-
ters and a model-form error were estimated from experimental rately [29]. Simple turbulent jets are better simulated using
measurements of turbulent kinetic energy k in a wind-tunnel simu- LEVM (rather than NLEVM) except for impinging jets [31]. For
lation of flow in a urban canyon. Gaussian process surrogates shock-boundary layer simulations, Loyau et al. [30] showed that

011001-2 / Vol. 4, MARCH 2018 Transactions of the ASME

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Cl ¼ Cl ðSÞ provided better results and cubic terms were not very the CEVM prediction AðS; X;ee Þc. The second term drives the
useful for those flows. Note that these jets did not have significant optimizer to set elements of c to zero as long as the first term does
curved streamlines or swirl. Studies in Ref. [32] found that not degrade. This allows us to “shrink” the CEVM, retaining only
QEVM, calibrated to different flows, resulted in a huge variability the terms that can be supported by the observational data and
in c1 ; c2 ; c3 and could not be used as a universal model. They for- prevents overfitting the CEVM model to limited Vobs data. The
mulated Eq. (1) instead. While it captures some of the anisotropy penalty k plays a crucial role and its optimal value is obtained
in normal turbulent stresses, it is not sufficient for flows with sig- via 11-fold cross-validation. Least absolute shrinkage and selec-
nificant heat transfer through boundaries and requires an extra tion operator [34] as implemented in R [35] via the glmnet
evolution equation for the anisotropy. Note that there are some package [36] is used to solve the linear inverse problem (Eq. (3)).
NLEVMs that do not adhere to the series structure in Eq. (1), e.g., The measured turbulent stresses Vobs and the CEVM-predicted
see Ref. [33]. ones V contain nondimensionalized normal and shear stresses
seij ¼ sij =Tij , where Tij ¼ jsobs
ij j is the mean of absolute values of
the measured turbulent stresses. Thus, a different normalization
2.4 Experimental Data. The wind-tunnel experiment that
constant is computed for s11 ; s22 , and s12.
provides us with calibration data is described in Refs. [6–8]. A
In Fig. 1 (top), we plot the number of coefficients retained in
schematic of the test section is shown in Fig. 1 of Ref. [4].
the CEVM model as k is increased; it is clear that as the penalty
The test section is 0.5 m long and has a square cross section of
increases, c becomes increasing sparse, retaining the most impor-
side 304.8 mm. The flow is from left to right, with variable Mach
tant coefficients ci required to minimize the k : k2 term. For
numbers from 0.5 to 0.8. The supersonic jet at Mach 3.73 is intro-
extreme k, the CEVM reduces to the intercept and linear terms. In
duced into the flow from a nozzle mounted on the floor of the tun-
Fig. 1 (bottom), we plot the results of the 11-fold cross-validation
nel test section. Particle image velocimetry is used to measure
velocities on the spanwise midplane (plane of symmetry) extend-
ing over a region of 200–400 mm downstream of the jet. For some
cases, measurements are also made on the crossplane, perpendicu-
lar to the flow direction, where the vorticity field is best captured.
Measurements inside a window W; 0  z  0:04 m; 0:03 m  y
 0:11 m, which captures one of the CVPs (see Fig. 1 in Ref. [4])
are used for calibration; it contains a 8  28 grid of observations.
Here, y is the vertical coordinate axis. We will call these locations
probes. A more comprehensive description of the experimental
setup is provided in Ref. [4].

3 Eddy Viscosity Model


The CEVM described in Eq. (1) has seven parameters and it is
unlikely that the information content of our experimental dataset
will allow their estimation in toto. Consequently, we simplify the
EVM. The measurements on the midplane are made at five down-
stream locations; each location contains 63 measurement points
(probes). Thus, the midplane measurements provide us with
5  63 ¼ 315 probe locations, where s11 ; s22 , and s12 measurements
exist. Further, particle image velocimetry measurements of the
mean flow on the midplane provide us with a well-resolved velocity
measurement, with experimental values of Sij and Xij computed at
the 315 midplane probes. This allows us to set up a linear problem
for CEVM-predicted turbulent stresses V ¼ AðS; X;ee Þc, where the
matrix AðS; X;ee Þ contains the strain-rate and vorticity terms in
Eq. (1) evaluated at the probes and c ¼ fc1 ; …; c7 g. AðS; X;ee Þ
has as many rows as the number of observations (nominally,
thrice the number of probes). Unfortunately, ee cannot be meas-
ured (and is typically available only inside an RANS simulation),
and consequently, we approximate it.
We assume that the rate of production of turbulent kinetic
energy is balanced by the dissipation, allowing us to set
e ¼ sij ð@=@xj ÞUi ; thereafter, ee can be computed using Eq. (2).
This assumption holds true only very approximately and at best
provides an order-of-magnitude estimate of ee . For many probes, ee
assumes (nonphysical) negative values and the probes have to be
eliminated. Further, probes in the nonturbulent freestream too
have to be excised, as both S and X are nearly zero there. In addi-
tion, we remove the probes that lie near the wall, in an attempt to
obtain a CEVM that is tuned for the CVP. Removing these meas-
urements leaves us with about 170 measurements that have large
S; X, and sij. Consolidating the measured, nonzero sij into Vobs ,
we pose the estimation of c as one of shrinkage

minimize kVobs  AðS; X;ee Þck22 þ kkck1 (3)


c Fig. 1 Top: Removal of CEVM coefficients ci as k is increased.
Bottom: Mean prediction error and 62 standard deviation
The first term in the objective function in Eq. (3) drives c to values bounds of prediction error, as a function of log(k). kmin (left ver-
that minimize the discrepancy between measurements Vobs and tical line) and k1se (right vertical line) are also shown.

ASCE-ASME Journal of Risk and Uncertainty in Engineering Systems, MARCH 2018, Vol. 4 / 011001-3
Part B: Mechanical Engineering

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


performed for increasing values of k. For each value of k, we implemented in the R [35] package FME [37]. The sufficiency of
obtain 11 mean-square errors (MSE); their average and 62 stand- samples is checked using the Raftery–Lewis method [38] imple-
ard deviation bounds are plotted. We identify kmin, the value for mented in the R package mcgibbsit [39]. A conjugate inverse
which we obtain the minimum average MSE (left vertical line); Gamma prior is used for P2 ðr2 Þ and is the same as in Ref. [4].
the digits on top of the plot show that this corresponds to a CEVM The samples of fC; r2 g drawn by the sampler are used to develop
with all the coefficients retained. However, we also obtain k1se , a histograms or PDFs using kernel density estimation. The MCMC
value of k for which the mean MSE is one standard deviation method may require Oð104 Þ samples to construct PðC; r2 jye Þ,
away from the one obtained via kmin. Further, this model requires each of which requires a 3D RANS simulation of JIC to provide
the retention of just one coefficient, c3, in the CEVM. This corre- ym ðCÞ. This is impractical and we replace the RANS simulator
sponds to the term with X2ij ; in a strongly vortical flow such as with a surrogate (also called a statistical emulator) to serve as a
JIC, this is not entirely surprising. computationally inexpensive proxy.
Thus, the NLEVM that we will use consists of two terms: the
linear term in Eq. (1) and one quadratic term involving X2ij . Note
that we do not show the value of c3 anywhere; we merely use least 4.2 Surrogate Models. A surrogate model, for the purposes
absolute shrinkage and selection operator to perform a data-driven of this study, is the functional dependence between vorticity xx
simplification of the CEVM. This is because the gross approxima- (or X23 ) in the streamwise direction as predicted by the RANS
tion used to obtain ee , for use in Eq. (3), makes the value of c3 so model, and the parameters C that cause it. As in Ref. [4], this
obtained untrustworthy. We will calibrate it, along with Ce2 and dependence is posed as a cubic polynomial in ðc3 ; Ce2 ; Ce1 Þ and
Ce1 , later in Sec. 5. learned by fitting (linear regression) to a training corpus of RANS
Note that it is possible that the quadratic term was retained runs.
because of its ability to correct the errors in the coarse approxima- We define a cuboidal parameter space C1 given by the follow-
tion of ee and consequently has little ability to favorably impact ing bounds: c3 2 ½0:1; 3:5; Ce2 2 ½1:7; 2:5, and Ce1 2 ½1:2; 1:7.
flow predictions. This will be investigated in Sec. 5 where we will The last two bounds were obtained from Ref. [26]. We draw 2744
calibrate c3, the factor that multiplies the quadratic term, using samples using a quasi-Monte Carlo sampler (Halton sequence)
flow observations. If it turns out that c3’s prior and posterior distri- and seed our RANS simulator SIGMA (see Ref. [4] for a descrip-
butions do not differ much, then the flow contains no information tion of the numerical scheme and test of mesh convergence; our
on the quadratic term in the NLEVM, and its retention was driven 3D finite volume meshes have about 10  106 cells). Most of the
by the coarse approximation of ee . If the converse is true, i.e., c3 is parameter combinations are nonphysical and only 222/2744 runs
constrained by flow observations, then the quadratic term merits survived, i.e., converged. These 222 are deemed to define a more
being included in an EVM. realistic parameter space C2 and we draw more samples from it to
provide a training corpus of sufficient size. We train an SVMC
with radial kernels (see Ref. [4] for a description) to segregate C2
4 Posing the Calibration Problem in C1 and identify 1500 C samples that lie in C2 . We run these
We seek to infer C ¼ ðc3 ; Ce2 ; Ce1 Þ from measurements of vor- samples again; 374/1500 samples survived. The 374 in turn define
ticity on the crossplane, obtained from the experiments reviewed C3 and we generate 1500 points from it using an SVMC. These
in Sec. 2. The method we adopt for doing so is described in Ref. are run and 1275/1500 runs survive. These 1275 samples are
[4], though there it was used to calibrate an LEVM-based RANS. deemed to be a parameter space C4 , where an RANS simulator
Reference [4] also contains a verification test of the Bayesian may be expected to run to completion. Note that we use the
parameter estimation method and software framework used in this CEVM in Ref. [5], with all the CEVM parameters except c3 set to
study. We provide a summary later. zero.
The runs using samples drawn from C3 do not necessarily pro-
vide flowfields that resemble physical ones; however, they do pro-
4.1 Bayesian Inverse Problem for Parameters. Let ye be a vide a vorticity field, however unrealistic, on the crossplane. We
vector (of length Np ) of experimental observations, measured at a set compare these vorticity fields with the experimental one, compute
of Np locations (probes). Let ym ðCÞ be model predictions of the the root-mean-square-error, and retain the best 20% of the runs.
same, produced by a parameter setting C. They are related by The parameter space occupied by these samples is the physically
ye ¼ ym ðCÞ þ e, where e is a combination of measurement and realistic part R. In Fig. 2 (top), crosses plot out C3 , whereas the
model-form error. We make a modeling assumption that the errors at filled circles denote R. The supporting information, which is
the probes are uncorrelated, independently and identically distributed available under the “Supplemental Data” tab for this paper on the
as a zero-mean Gaussian, i.e., e ¼ fei g; ei  N ð0; r2 Þ. r2 thus pro- ASME Digital Collection, contains projections on the relevant 2D
vides a crude measure of the model—data misfit after calibration. planes.
Let PðC; r2 jye Þ be the JPDF of the parameters and the model— The 0:2  1275 ¼ 255 RANS runs constituting R are used to
data misfit that we seek; it is also called the “posterior” density. construct surrogate models. We first identify the subset of 8 
Let P1 ðCÞ and P2 ðr2 Þ be our prior belief regarding the distribu- 28 ¼ 224 probes in W, where vorticity is large (see Ref. [4] for
tion of C and r2 . The likelihood of observing ye , given a parame- details). At these probes, we impute a cubic polynomial in
ter setting C; Lðye jCÞ, is given by ðc3 ; Ce2 ; Ce1 Þ to model vorticity and fit it via linear regression. We
  use Akaike information criterion to simplify the model; for most
 2
 1 jjye  ym ðCÞjj22 probes, the models reduce to quadratic polynomials in
L ye jC; r / Np exp 
r 2r2 ðc3 ; Ce2 ; Ce1 Þ. The surrogate models so obtained are not very
robust (small changes in the training corpus could lead to very dif-
and by Bayes’ theorem ferent surrogates). Consequently, as in Ref. [4], we perform 50
    rounds of repeated random subsampling (a form of cross-valida-
P C; r2 jye / L ye jC; r2 P1 ðCÞ P2 ðr2 Þ tion) to compute robust surrogates. Surrogates whose testing set
  (TS) errors are less than 10% and where the ratio of learning and
1 jjy  ym ðCÞjj22
/ Np exp  e testing set errors deviates from 1 by less than 15% are deemed fit
r 2r2 for calibration purposes. Thus, accurate surrogate models could
 P1 ðCÞ P2 ðr Þ
2
(4) not be constructed for all the probes with large vorticity. We plot
the testing set errors and the ratios in Fig. 2 of the supporting
This inverse problem is solved via MCMC sampling. Identical to information, which is available under the “Supplemental Data”
Ref. [4], we use delayed rejection adaptive metropolis as tab for this paper on the ASME Digital Collection. In Fig. 2, we

011001-4 / Vol. 4, MARCH 2018 Transactions of the ASME

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


function is approximated by the SVMC with radial kernels, as
described in Ref. [4]. We test the SVMC’s accuracy via repeated
random subsampling cross-validation. We randomly split the
training data into learning set (LS)/testing set (TS) Pairs, with
85% of the training data lying in the LS. The SVMC is trained on
the LS and is tested on the TS to obtain the misclassification rate.
This is performed 50 times, with different LS/TS pairs, and an
average misclassification rate is computed. We find it to be 6.5%.
The SVMC is used to compute fm ðCÞ, the approximation of
fðCÞ, at arbitrary C, and identifies whether C 2 R. Thereafter, the
computation of P1 ðCÞ is straightforward via Eq. (5).

5 Results
The Bayesian inverse problem in Eq. (4) requires 80,000
MCMC steps to converge. These samples are thinned by a factor
of 5 (i.e., 1 in 5 samples is retained) and used to describe the pos-
terior density. In Fig. 3, we plot the marginalized PDFs for all the
parameters as well as r2 . The nominal values of the parameters
are also plotted. We see that the nominal value for Ce1 approxi-
mately agrees with the maximum a posteriori value (MAP, peak
of the PDF), but for the rest of the parameters the MAP values
and the nominal ones are quite different. In fact, the nominal val-
ues are in the tails of the prior distribution, indicating their subop-
timality. The PDFs of all the three parameters are bimodal,
another sign of the complexity of the posterior density. Further,
the marginalized PDF for c3 is quite different from its prior, indi-
cating that the observations contain information on c3. In addition,
we plot the model-data misfit, which has units of the calibration
variable/observable, the vorticity on the crossplane. The variations
of the magnitude of vorticity predicted by the best set of parame-
ters (Copt ; given later), i.e., its 5th, 25th, 50th, 75th, and
95th, percentiles, are f0:007; 0:039; 0:76; 2:4; 4:1g  103 s1 . The
experimental counterparts are f0:0093; 0:069; 0:184; 0:89; 3:8g 
103 s1. The sign of the vorticity is negative. Comparing r with
the 95th percentile of the vorticity magnitude, we see that r is
quite large (about 20%). Note that a slight offset in two sharply
defined vorticity fields (such as the CVP) can result in a very large
r. A better comparison is a plot of the two vorticity fields; this is
discussed in Fig. 4. In Fig. 3 in the supporting information, which
Fig. 2 Top: Plot of the physically realistic part of the parameter
space, R (filled circles) along with the region where the RANS
simulator runs without crashing C3 (crosses). Bottom: The
experimental vorticity field as a flood plot, with locations with
large vorticity (䊊) and the subset of probes with accurate surro-
gate models (1).

plot the 90 probes inside W with large vorticity. A subset of 50/90


probes, plotted with “plus” signs, are the ones where we could
construct accurate surrogate models. Thus, Np ¼ 50.
Altogether, about 6000 3D RANS JIC simulations were con-
ducted. Each simulation takes approximately 12,000 core-hours
on a PowerPC A2 processor.

4.3 Constructing an Informative Prior P1 ðCÞ. The surro-


gate models constructed in Sec. 4.2 are only valid inside R  C4 ,
and consequently, the posterior density (Eq. (4)) must be solved
within that constraint. We employ the prior belief

1 for C 2 R
P1 ðCÞ ¼ (5)
0 otherwise

We enforce this prior belief using an SVMC.


Fig. 3 Marginalized PDFs of c3 ; Ce 2; Ce 1, and r2 . The dashed
Fundamentally, we seek a binary function fðCÞ in ðc3 ; Ce2 ; Ce1 Þ line is the prior density due to R and the solid line denotes the
space that could be used to decide whether a parameter combina- posterior density. The vertical lines are the nominal values of
tion C 2 R. We set fðCÞ ¼ 1 at the 255/1275 that defines R and the parameters. The figure at the bottom right shows magnitude
fðCÞ ¼ 1 elsewhere. Then, fðCÞ ¼ 0 denotes @R, the demarca- of the data—model misfit. As a comparison to r, the 95th per-
tion of the physically realistic part of the parameter space. This centile of the (experimental) vorticity magnitude is 3:83103 s21.

ASCE-ASME Journal of Risk and Uncertainty in Engineering Systems, MARCH 2018, Vol. 4 / 011001-5
Part B: Mechanical Engineering

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 4 Plots of simulated streamwise vorticity field (as a flood plot) with contours of experimental vorticity overlaid. Left:
Simulations using RANS-LEVM driven by nominal parameters. Middle: RANS–CEVM, driven by the nominal parameters in
Ref. [5]. Right: Predictions using Copt . The improvement is stark. Note that the scales of the vertical and horizontal axes
are different.

is available under the “Supplemental Data” tab for this paper on experimental measurements. Most of e is due to model-form error
the ASME Digital Collection, we plot 2D marginal densities for (as opposed to measurement uncertainty) and shows the impact of
the parameters. We see that they are complex and strong correla- approximate modeling of turbulent processes in the k  e model.
tions exist between all and especially for Ce2 and Ce1 . The two Next, we investigate the response of the RANS–CEVM simula-
modes in the density are also clearly evident. tor (RANS with an embedded CEVM with seven parameters) to
First, we check the accuracy of the calibration via a posterior the 100 ðc3 ; Ce2 ; Ce1 Þ samples from the posterior density. All the
predictive test. We randomly select 100 ðc3 ; Ce2 ; Ce1 ; r2 Þ samples CEVM parameters except c3 are set to zero. We perform the simu-
from those picked by the MCMC sampler and predict the vorticity lations and summarize the streamwise vorticity field inside W by
at the probes with accurate surrogate models. Note that these pre- the circulation, the vorticity field’s centroid in the crossplane, and
dictions also contain a realization of the model-data mismatch e, the radius of gyration. These provide a point-vortex approxima-
corresponding to N ð0; r2 Þ. These predictions are then normalized tion of the vorticity field, and we refer to them as the “point-
by their experimental counterparts. We use this ensemble to com- vortex metrics” (PVM). These metrics are normalized by their
pute the 5th, 50th, and 95th percentiles of the prediction at the experimental counterparts and plotted in Fig. 6. No e is added to
probes. In Fig. 5, we plot the normalized vorticity at the probes them. The outlines of the boxes are the first and third quartiles of
where they are available. The filled 䊊 is the median prediction and the predictions, whereas the horizontal line is the median. The
the error bars span the 5th–95th percentile range. The horizontal whiskers denote the limits beyond which we have outliers. The
line at 1 denotes the experimental data. It is clear that the experi- small variability of the PVM is a reflection of the short error bars
mental values are consistently bracketed by the error bars. Note in Fig. 4 in the supporting information, which is available under
that much of the variability (the width of the error bars) is due to the “Supplemental Data” tab for this paper on the ASME Digital
e; Fig. 4 in supporting information, which is available under the Collection. The horizontal line at y ¼ 1 indicates the experimental
“Supplemental Data” tab for this paper on the ASME Digital Col- value. The open 䊊 are predictions using RANS–CEVM, seeded
lection, contains the same plot, but without e (we call this the with c1 ; …; c7 from Ref. [5], while the filled 䉫 are predictions
“pushed-forward posterior”). As is clear, the shorter error bars in with RANS-LEVM, seeded with the nominal values fCl ;Ce2 ;Ce1 g
the pushed-forward posterior tests sometimes do not bracket the ¼ f0:09;1:92;1:44g. We see that while using nominal values of
the parameters, LEVM is better than CEVM. This is somewhat
unexpected as the severe bending of the jet and the CVP should
have played to the strengths of CEVM. However, selecting one

Fig. 5 Results of the posterior predictive test using 100 sam- Fig. 6 Box-and-whisker plots of the posterior samples’ runs
ples. We plot the predicted vorticity normalized by the meas- with RANS–CEVM, for the PVM. The horizontal line denotes
ured value at the probes with accurate surrogate models. The experimental results. The open 䊊 are prediction using nominal
horizontal line indicates the experimental measurement. The CEVM parameters [5], whereas the filled 䉫 are predictions
error bars span the 5th–95th percentile range and the filled 䊊 using an LEVM with nominal parameters. Model predictions are
are the median prediction. normalized by their experimental counterparts.

011001-6 / Vol. 4, MARCH 2018 Transactions of the ASME

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


quadratic term in the CEVM and calibrating it results in a ensemble’s predictions using the mean. These plots are shown in
RANS simulator that is clearly superior to either of the two unca- Fig. 7 at three locations 200, 300, and 400 mm downstream of the
librated models. This seems to indicate that the inferiority of jet. The top row contains the velocity deficit and the lower row
CEVM may be due to the values of c1 ;…;c7 in Ref. [5] being the vertical velocity. Predictions with the uncalibrated parameters
unsuitable for transonic JIC interactions. Postcalibration, the (LEVM or CEVM), i.e., the dotted and dashed lines, lead to a jet
PVM have errors less than 20%, whereas predictions with nominal that sits too far above the true (experimental) jet; this is clear in
parameter values lead to 30–40% error. The ensemble of both the deficit and vertical velocity plots. Further, the vertical
100 runs also allowed us to obtain an optimal value of velocity is overpredicted (bottom row), indicating a CVP that is
C; Copt ¼ ðc3 ;Ce2 ;Ce1 Þopt ¼ f0:169;2:464;1:496g, which yielded too strong. This is in agreement with the circulation being over-
PVM that were closest to 1. We will refer to this as our “best predicted (before calibration) as shown in Fig. 6. It is also clear
case.” that the LEVM results agree with the experimental data better
In Fig. 4, we investigate the vorticity field on the crossplane than the CEVM ones before calibration. However, the improve-
further. We plot the streamwise vorticity field predicted by ment in agreement, after calibrating ðc3 ; Ce2 ; Ce1 Þ, is quite sub-
RANS-LEVM (left) and RANS–CEVM (right) using their nomi- stantial. The mean streamwise velocity deficit and vertical
nal values. Overlaid on them is the experimental vorticity field, velocity are quite close to the experimental data, with the agree-
plotted as contours. As is clear, the agreement is not very good; ment in vertical velocity being superior, compared to the stream-
further, the vorticity field produced by the CEVM is substantially wise velocity deficit. In addition, the agreement in vertical
worse than LEVM. On the right is the vorticity field predicted by velocity improves as we proceed downstream. Further, the predic-
Copt . The improvement is substantial, further reinforcing the tion using Copt (þ symbols in Fig. 7) is almost identical to the
unsuitability of RANS–CEVM (with nominal parameter values) ensemble mean.
for JIC. Finally, we check if the calibration performed using a M ¼ 0.8
The improvement in vorticity field predictions after the estima- crossflow is applicable at other flow conditions. We use the 100
tion of ðc3 ; Ce2 ; Ce1 Þ is somewhat expected since that was the cali- ðc3 ; Ce2 ; Ce1 Þ samples to simulate a JIC interaction with M ¼ 0.7
bration variable. We now investigate whether calibration crossflow. The experimental dataset [6] contains measurements of
improves the entire flowfield. In order to do so, we compare the velocity on the midplane for a M ¼ 0.7 interaction, but we do not
streamwise velocity deficit (ðUmax  uÞ=U1 ) and normalized ver- have any crossplane measurements. Consequently, we plot the
tical velocity (v=U1 ) profiles as a function of the normalized equivalent of Fig. 7 for the M ¼ 0.7 interaction in Fig. 8. The find-
height y=Dj with experimental measurements. Here, U1 is the ings of Fig. 7 hold here too—LEVM (nominal parameter values)
freestream velocity (286 m/s) and Dj is the jet diameter provides better predictions than CEVM (nominal parameter val-
(9.83 mm). Umax is the maximum velocity in test section and is ues) and the calibration produced with M ¼ 0.8 crossflow leads to
slightly higher than U1 due to addition of extra mass by the jet. a remarkably good prediction (via the ensemble mean) of the
We use the midplane flowfields simulated using the 100 M ¼ 0.7 experimental results, especially for the vertical velocity
ðc3 ; Ce2 ; Ce1 Þ samples described earlier to compute the streamwise (bottom row). Thus, there is some degree of robustness in the cali-
velocity deficit and the vertical velocity. We summarize the bration of ðc3 ; Ce2 ; Ce1 Þ.

Fig. 7 Plots of streamwise velocity deficit (top row) and vertical velocity (bottom row) at three
locations 200, 300, and 400 mm downstream of the jet. Experimental data are plotted using sym-
bols, LEVM (nominal) using the dotted line, the CEVM (nominal) using the dashed line, and the
ensemble mean of 100 samples from the posterior using the solid line. The 1 symbols are the
predictions using Copt . The crossflow is M 5 0.8.

ASCE-ASME Journal of Risk and Uncertainty in Engineering Systems, MARCH 2018, Vol. 4 / 011001-7
Part B: Mechanical Engineering

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 8 Plots of streamwise velocity deficit (top row) and vertical velocity (bottom row) at three
locations 200, 300, and 400 mm downstream of the jet. Experimental data are plotted using sym-
bols, LEVM (nominal) using the dotted line, the CEVM (nominal) using the dashed line, and the
ensemble mean of 100 samples from the posterior using the solid line. The 1 symbols are the
predictions using Copt . The crossflow is M 5 0.7.

5.1 Discussion. Predictive errors in RANS simulations are due The impact of adding the X2ij term to the EVM is clearly shown
to, among other causes, the shortcomings of the EVM and the in Figs. 7 and 8. As the CVP evolves downstream, it widens, i.e.,
approximate closure relationships in the evolution equations for k and a larger fraction of the flowfield contains large vorticity. The vor-
e. In the results mentioned earlier, we see that simply sensitizing the ticity field modulates sij. As is seen in the two figures, agreement
turbulent stresses to strain-rate, vorticity, streamline curvature, etc., of the vertical velocity clearly improves downstream. This
serves little purpose unless the parameters accompanying the model dependence was not seen in our previous study [4]. The results
are also properly calibrated. The lack of free parameters to calibrate mentioned earlier also show that the calibration of ðc3 ; Ce2 ; Ce1 Þ
thus renders the simple LEVM more predictive for JIC than CEVM improves the predictive skill of the entire flowfield and not just
when the nominal parameter values are used. Our EVM which has streamwise vorticity. Further, it is robust across changes in the
only one quadratic term in CEVM “turned on” (the one multiplying crossflow Mach number. This is in line with what was observed in
X2ij ) provides superior results, but only after calibration. Further, since our previous paper which did not address enrichment of the
we calibrate Ce2 and Ce1 (which appear in the evolution equation of EVM [4].
e) in addition to c3, it is not clear whether our calibrated EVM or the
e-equation is responsible for this improvement.
In our previous work [4], we had addressed the improvement of 6 Conclusions
an RANS equation (with LEVM), calibrated to the same experi- In this study, we investigate whether enriching a standard linear
mental data. There we estimated ðCl ; Ce2 ; Ce1 Þ and obtained an eddy viscosity model with higher-order terms could reduce the
improvement in the predictive skill similar to the one seen in this model-form errors in RANS simulations of jet-in-crossflow inter-
paper (or perhaps even slightly better). Our previous paper actions. In particular, we choose a cubic eddy viscosity model [5].
allowed us to directly control the dependence of turbulent stresses We find that simulations using the linear eddy viscosity model are
on strain-rate (via the LEVM) during calibration, whereas in the actually superior to those using the cubic model, as long as both
present study, we control its dependence on vorticity (via our are seeded with parameters obtained from literature, i.e., they use
NLEVM). However, both the studies provide a similar degree of nominal parameter values. This highlights the crucial role played
improvement in predictive skill, inviting the speculation that it by the eddy viscosity model’s parameters and the importance of
may be ðCe2 ; Ce1 Þ, the parameters in the e-equation, which play calibrating them to relevant experimental data. Thus, the type of
the dominant role in drawing the flowfield closer to experimental model enrichment that may be possible is inextricably linked to
measurements. Further, the PDFs for Ce2 and Ce1 , drawn from the available measurements. We devise a statistical procedure,
these two studies (see Figs. 3 and 5 in Ref. [4]), show that their based on shrinkage, to discover a high-order eddy viscosity model
MAP values are somewhat similar—for Ce2 they are far greater in a purely data-driven manner. We find that our experimental
than the nominal value of 1.92, whereas Ce1 is quite close to the dataset can support an additional term in the eddy viscosity model
nominal one of 1.44. The insensitivity of these crucial parameters that is quadratic in vorticity. Given the strongly vortical nature of
to the form of the EVM tends to argue that the impact of c3 (or Cl the jet-in-crossflow interaction, this is not surprising.
in LEVM) may be of secondary importance, and the equations for We embed the new eddy viscosity model in an RANS simulator
k and e may be more important for improving the predictive skill and calibrate its parameter (c3) along with two other parameters,
of RANS for JIC interactions. Ce2 and Ce1 , that appear in the evolution equation for e, the

011001-8 / Vol. 4, MARCH 2018 Transactions of the ASME

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


dissipation rate of the turbulent kinetic energy. We use vorticity [3] Durbin, P. A., 1995, “Separated Flow Computations Using the k  e  v2
measurements on the crossplane of a jet-in-crossflow interaction as Model,” AIAA J., 33(4), pp. 659–664.
[4] Ray, J., Lefantzi, S., Arunajatesan, S., and Dechant, L., 2016, “Bayesian Param-
the calibration observable. The Mach number M of the crossflow is eter Estimation of a k-Epsilon Model for Accurate Jet-in-Crossflow Simu-
0.8. The calibration is Bayesian and we compute a joint PDF for lations,” Am. Inst. Aeronaut. Astronaut. J., 54(8), pp. 2432–2448.
the parameters being estimated. We use surrogate models to cir- [5] Craft, T. J., Launder, B. E., and Suga, K., 1996, “Development and Application
cumvent the enormous computational expense of Bayesian solu- of a Cubic Eddy-Viscosity Model of Turbulence,” Int. J. Heat Fluid Flow,
17(2), pp. 108–115.
tions of inverse problems. We also use informative priors to excise [6] Beresh, S. J., Henfling, J. F., Erven, R. J., and Spillers, R. W., 2005, “Penetration
nonphysical parts of the parameter space. We find that calibration of a Transverse Supersonic Jet Into a Subsonic Compressible Crossflow,” AIAA
immensely improves the predictive skill of the simulator. This J., 43(2), pp. 379–389.
improvement is seen in the entire flowfield (and not just vorticity [7] Beresh, S. J., Henfling, J. F., Erven, R. J., and Spillers, R. W., 2005, “Turbulent
Characteristics of a Transverse Supersonic Jet in a Subsonic Compressible
on the crossplane) and holds even when the calibrated parameters Crossflow,” AIAA J., 43(11), pp. 2385–2394.
are used to predict a jet-in-crossflow interaction with a M ¼ 0.7 [8] Beresh, S. J., Henfling, J. F., Erven, R. J., and Spillers, R. W., 2006,
crossflow. Thus, the calibration has some degree of robustness. “Crossplane Velocimetry of a Transverse Supersonic Jet in a Transonic Cross-
The joint calibration of ðc3 ; Ce2 ; Ce1 Þ in this study, taken in iso- flow,” Am. Inst. Aeronaut. Astronaut. J., 44(12), pp. 3051–3061.
[9] Arunajatesan, S., 2012, “Evaluation of Two-Equation RANS Models for Simu-
lation, does not allow us to ascribe the improvement in predictive lation of Jet-in-Crossflow Problems,” AIAA Paper No. 2012-1199.
skill to the addition of the X2ij term in the eddy viscosity model. [10] Chai, X., and Mahesh, K., 2011, “Simulations of High Speed Turbulent Jets in
However, in a previous study [4], we had calibrated an RANS Crossflow,” AIAA Paper No. 2010-4603.
simulator with a linear eddy viscosity model to the same experi- [11] Genin, F., and Menon, S., 2010, “Dynamics of Sonic Jet Injection Into Super-
sonic Crossflow,” J. Turbul., 11, pp. 1–13.
mental dataset. The parameters calibrated were ðCl ; Ce2 ; Ce1 Þ. The [12] Brinckman, K. W., Calhoon, W. H., and Dash, S. M., 2007, “Scalar Fluctuation
improvements in predictive skill and the robustness of the calibra- Modeling for High-Speed Aeropropulsive Flows,” AIAA J., 45(5), pp.
tion are very similar to the ones obtained in this study. Further, 1036–1046.
the PDFs for Ce2 and Ce1 also show some agreement. Thus, the [13] Papp, J. L., and Dash, S. M., 2001, “Turbulence Model Unification and Assess-
ment for High-Speed Aeropropulsive Flows,” AIAA Paper No. 2001-0880.
addition of a X2ij term in the eddy viscosity model had a rather [14] Edeling, W. N., Cinnella, P., Dwight, R. P., and Bijl, H., 2014, “Bayesian Esti-
muted effect. This was unexpected given the strongly vortical mates of Parameter Variability in k  e Turbulence Model,” J. Comput. Phys.,
nature of the jet-in-crossflow interaction, where the primary 258, pp. 73–94.
coherent structure is a CVP. It supports the notion that the modifi- [15] Cheung, S. H., Oliver, T. A., Prudencio, E. E., Prudhomme, S., and Moser, R.
D., 2011, “Bayesian Uncertainty Analysis With Applications to Turbulence
cation of the closure models (in the e-equation) by Ce2 and Ce1 Modeling,” Reliab. Eng. Syst. Saf., 96(9), pp. 1137–1149.
plays a significant, if not dominant, role in improving the predic- [16] Gilks, W. R., Richardson, S., and Spiegelhalter, D. J., eds., 1996, Markov Chain
tive skill of RANS models. It also implies that the closure models Monte Carlo in Practice, Chapman & Hall/CRC Press, Boca Raton, FL.
may be equally important contributors to the model-form errors in [17] Duraisamy, K., Zhang, Z. J., and Singh, A. P., 2015, “New Approaches in Tur-
bulence and Transition Modeling Using Data-Driven Techniques,” AIAA Paper
RANS jet-in-crossflow simulations, along with the eddy viscosity No. 2015-1284.
model. [18] Zhang, Z. J., and Duraisamy, K., 2015, “Machine Learning Methods for Data-
While the enrichment of the eddy viscosity model with a quad- Driven Turbulence Modeling,” AIAA Paper No. 2015-2460.
ratic vorticity term may not have caused large global changes in [19] Zhang, Z., Duraisamy, K., and Gumerov, N. A., 2015, “Efficient Multiscale Gaus-
sian Process Regression Using Hierarchical Clustering,” e-print arXiv:1511.02258.
flowfield, it may have intense local effects, e.g., inside the [20] Pratap Singh, A., Medida, S., and Duraisamy, K., 2016, “Machine Learning-
counter-rotating vortex pair. This could be evidenced in an RANS Augmented Predictive Modeling of Turbulent Separated Flows Over Airfoils,”
simulation of a jet-in-crossflow interaction where vorticity meas- e-print arXiv:1608.03990.
urements on the crossplane are available. This particular measure- [21] Singh, A. P., and Duraisamy, K., 2016, “Using Field Inversion to Quantify
Functional Errors in Turbulence Closures,” Phys. Fluids, 28(4), p. 045110.
ment is not available in our M ¼ 0.7 dataset, and the investigation [22] Xiao, H., Wu, J.-L., Wang, J.-X., Sun, R., and Roy, C., 2016, “Quantifying and
is left for future work. Reducing Model-Form Uncertainties in Reynolds-Averaged Navier–Stokes
Simulations: A Data-Driven, Physics-Informed Bayesian Approach,” J. Com-
put. Phys., 324, pp. 115–136.
[23] Wang, J.-X., Wu, J.-L., and Xiao, H., 2016, “A Physics Informed Machine
Acknowledgment Learning Approach for Reconstructing Reynolds Stress Modeling Discrepan-
This work was supported by Sandia National Laboratories’ cies Based on DNS Data,” e-print arXiv:1606.07987.
[24] Wu, J.-L., Wang, J.-X., Xiao, H., and Ling, J., 2016, “Physics-Informed
Advanced Scientific Computing (ASC) Verification and Valida- Machine Learning for Predictive Turbulence Modeling: A Priori Assessment of
tion program. Sandia National Laboratories is a multi-program Prediction Confidence,” e-print arXiv:1607.04563.
laboratory managed and operated by Sandia Corporation, a wholly [25] Weatheritt, J., and Sandberg, R., 2016, “A Novel Evolutionary Algorithm
owned subsidiary of Lockheed Martin Corporation, for the U.S. Applied to Algebraic Modifications of the RANS Stress–Strain Relationship,”
J. Comput. Phys., 325, pp. 22–37.
Department of Energy’s National Nuclear Security Administration [26] Guillas, S., Glover, N., and Malki-Epshtein, L., 2014, “Bayesian Calibration of
under Contract No. DE-AC04-94AL85000. We thank Lawrence the Constants of the k  e Turbulence Model for a CFD Model of Street Canyon
Livermore National Laboratory, Livermore, CA, for computer Flow,” Comput. Methods Appl. Mech. Eng., 279, pp. 536–553.
time on the Sequoia supercomputer, a National User Facility. [27] Haario, H., Laine, M., Mira, A., and Saksman, E., 2006, “DRAM-Efficient
Adaptive MCMC,” Stat. Comput., 16(4), pp. 339–354.
[28] Parish, E., and Duraisamy, K., 2015, “Quantification of Turbulence Modeling
Nomenclature Uncertainties Using Full Field Inversions,” AIAA Paper No. 2015-2459.
[29] Bazdidi-Tehrani, F., Mohammadi-Ahmar, A., Kianamsouri, M., and Jadidi, M.,
C¼ parameters to be calibrated 2015, “Investigation of Various Non-Linear Eddy Viscosity Turbulence Models
Ce2 ; Ce1 ¼parameters in the k  e model, to be calibrated for Simulating Flow and Pollutant Dispersion on and Around a Cubical Model
J¼ jet-to-crossflow momentum ratio Building,” Build. Simul., 8(2), pp. 149–166.
[30] Loyau, H., Batten, P., and Leschziner, M. A., 1998, “Modelling Shock/Bound-
R¼ the physically relevant part of the C parameter space ary-Layer Interaction With Nonlinear Eddy-Viscosity Closures,” Flow Turbul.
S¼ strain-rate tensor Combust., 60(3), pp. 257–282.
N ðl; r2 Þ ¼normal distribution with mean l and standard devia- [31] Balabel, A., and El-Askary, W. A., 2011, “On the Performance of Linear and
tion r Nonlinear k–e Turbulence Models in Various Jet Flow Applications,” Eur. J.
Mech. B/Fluids, 30(3), pp. 325–340.
X ¼ vorticity tensor [32] Craft, T. J., Launder, B. E., and Suga, K., 1997, “Prediction of Turbulent Tran-
sitional Phenomena With a Nonlinear Eddy-Viscosity Model,” Int. J. Heat Fluid
References Flow, 18(1), pp. 15–28.
[1] Poroseva, S., and Iaccarino, G., 2001, “Simulating Separated Flows Using the [33] Gatski, T. B., and Speziale, C. G., 1992, “On Explicit Algebraic Stress Models
k  e Model,” Center for Turbulence Research, Stanford University, Stanford, for Complex Turbulent Flows,” Institute for Computer Application in Science
CA, pp. 375–383. and Engineering, NASA Langley Research Center, Hampton, VA, Technical
[2] Sj€ogren, T., and Johansson, A. V., 2000, “Development and Calibration of Report No. 92-58.
Algebraic Nonlinear Models for Terms in the Reynolds Stress Transport Equa- [34] Hastie, T., Tibshirani, R., and Friedman, J., 2008, The Elements of Statistical
tion,” Phys. Fluids, 12(6), pp. 1554–1572. Learning (Springer Series in Statistics), 2 ed., Springer, New York.

ASCE-ASME Journal of Risk and Uncertainty in Engineering Systems, MARCH 2018, Vol. 4 / 011001-9
Part B: Mechanical Engineering

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


[35] R Core Team, 2012, “R: A Language and Environment for Statistical [38] Raftery, A., and Lewis, S. M., 1996, “Implementing MCMC,” Markov Chain
Computing,” R Foundation for Statistical Computing, Vienna, Austria. Monte Carlo in Practice, W. R. Gilks, S. Richardson, and D. J. Spiegelhalter,
[36] Friedman, J., Hastie, T., and Tibshirani, R., 2010, “Regularization Paths for eds., Chapman & Hall/CRC Press, Boca Raton, FL, pp. 115–130.
Generalized Linear Models Via Coordinate Descent,” J. Stat. Software, 33(1), [39] Warnes, G. R., and Burrows, R., 2011, “Mcgibbsit: Warnes and Raftery’s
pp. 1–22. MCGibbsit MCMC Diagnostic. R Package Version 1.0.8,” accessed Aug. 15,
[37] Soetaert, K., and Petzoldt, T., 2010, “Inverse Modelling, Sensitivity and Monte 2017, https://artax.karlin.mff.cuni.cz/r-help/library/mcgibbsit/html/mcgibbsit.
Carlo Analysis in R Using Package FME,” J. Stat. Software, 33(3), pp. 1–28. html

011001-10 / Vol. 4, MARCH 2018 Transactions of the ASME

Downloaded From: http://risk.asmedigitalcollection.asme.org/ on 10/21/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like