You are on page 1of 17

AIAA JOURNAL

Robust Bayesian Calibration of a k − ϵ Model for Compressible


Jet-in-Crossflow Simulations

Jaideep Ray∗
Sandia National Laboratories, Livermore, California 94550-0969
Lawrence Dechant† and Sophia Lefantzi‡
Sandia National Laboratories, Albuquerque, New Mexico 87185-5800
Julia Ling§
Citrine Informatics, Redwood City, California 94063
and
Srinivasan Arunajatesan¶
Sandia National Laboratories, Albuquerque, New Mexico 87185-5800
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

DOI: 10.2514/1.J057204
Compressible jet-in-crossflow interactions are difficult to simulate accurately using Reynolds-averaged Navier–
Stokes (RANS) models. This could be due to simplifications inherent in RANS or the use of inappropriate RANS
constants estimated by fitting to experiments of simple or canonical flows. Our previous work on Bayesian calibration
of a k − ϵ model to experimental data had led to a weak hypothesis that inaccurate simulations could be due to
inappropriate constants more than model-form inadequacies of RANS. In this work, Bayesian calibration of k − ϵ
constants to a set of experiments that span a range of Mach numbers and jet strengths has been performed. The
variation of the calibrated constants has been checked to assess the degree to which parametric estimates compensate
for RANS’s model-form errors. An analytical model of jet-in-crossflow interactions has also been developed, and
estimates of k − ϵ constants that are free of any conflation of parametric and RANS’s model-form uncertainties
have been obtained. It has been found that the analytical k − ϵ constants provide mean-flow predictions that are
similar to those provided by the calibrated constants. Further, both of them provide predictions that are far closer to
experimental measurements than those computed using “nominal” values of these constants simply obtained from the
literature. It can be concluded that the lack of predictive skill of RANS jet-in-crossflow simulations is mostly due to
parametric inadequacies, and our analytical estimates may provide a simple way of obtaining predictive compressible
jet-in-crossflow simulations.

Nomenclature M = crossflow Mach number


C = parameters in the k − ϵ RANS model to be N μ; σ 2  = normal distribution with mean μ and standard
calibrated deviation σ
Ca = analytical estimate of C TD = training data
Cnom = nominal values of C udef = velocity deficit in the streamwise direction
Copt = optimal value of C vnorm = normalized vertical velocity
x = streamwise distance
CVP = counterrotating vortex pair
δm = structural error; difference between exper-
Cμ = A parameter in the eddy-viscosity submodel
imental data and calibrated surrogate model
in a k − ϵ RANS model
predictions
Cϵ2 , Cϵ1 = parameters in the equation of the evolution of
σ = model–data mismatch is represented as
ϵ in a k − ϵ RANS model
N 0; σ 2 
d = approximation error in the surrogate model
ξ = normalized radial distance, r∕lx
fξ, gξ, hξ = normalized radial profiles for streamwise
velocity, turbulent kinetic energy, and
dissipation
J = jet-to-crossflow momentum ratio I. Introduction
JPDF = joint probability density function
lx = jet length-scale at streamwise location x T URBULENT jet-in-crossflow (JIC) interactions occur in many
natural and engineering situations [1]. One such occurrence, in
aerodynamics, is the stabilization of aerodynamic bodies using spin
rockets mounted perpendicular to the direction of flight. Here the
Presented as Paper 2017-4166 at the AIAA Aviation and Aeronautics
Forum and Exposition, Denver, CO, 25–29 June 2017; received 10 February
interaction of the rocket exhaust with the freestream constitutes
2018; revision received 29 June 2018; accepted for publication 5 July 2018; a JIC interaction. It is known that the exhaust can interact with the
published online 28 September 2018. Copyright © 2018 by the American control surfaces of the aerodynamic body, modifying the pressure
Institute of Aeronautics and Astronautics, Inc. All rights reserved. All requests distribution and interfering with the moments experienced by the
for copying and permission to reprint should be submitted to CCC at body [2]. Experimental investigations of the JIC interactions [3–6]
www.copyright.com; employ the ISSN 0001-1452 (print) or 1533-385X show that the exhaust (the jet) rolls into a counterrotating vortex
(online) to initiate your request. See also AIAA Rights and Permissions pair (CVP), which then interacts with the fin/control surface [7]; a
www.aiaa.org/randp. vortex-dynamical model of this interaction has also been proposed.
*Technical Staff, Extreme Scale Data Science and Analytics, MS 9152. However, numerical simulations using k − ϵ and k − ω RANS
Senior Member AIAA.

Technical Staff, Aerosciences Department, MS 0825. Senior Member models could only reproduce the experimental measurements
AIAA. qualitatively [8].

Manager, Fluid and Reactive Processes, MS 0828. Senior Member AIAA. The lack of predictive skill of the RANS simulations could be due
§
Manager, Data Science. to two causes. First, RANS models contain constants whose values

Manager, Aerosciences Department, MS 0825. Senior Member AIAA. are estimated by fitting to experimental data gathered from simple
Article in Advance / 1
2 Article in Advance / RAY ET AL.

flows. These flows have little in common with JIC interactions. analytically, and evidence from other researchers that values of C
Further, RANS parameters are often tuned for different classes of other than Cnom might be predictive for JIC interactions. In Sec. III,
flows; see Ref. [9] for a discussion. Second, two-equation RANS we describe the inverse modeling methodology used to compute the
models contain gross simplifications of turbulent processes [10], JPDFs from experimental data. In Sec. IV, we develop the analytical
giving rise to model-form errors. It is unclear which of the two causes model. In Sec. V we present our results, which are then discussed in
contributes more to the inaccuracies in JIC simulations. In one of Sec. VI. In Sec. VII, we present our conclusions.
our previous papers [9], we investigated the first possible cause of
errors, that is, whether JIC simulations could be improved by using
better RANS parameters. We developed a Bayesian technique for II. Background
calibrating three k − ϵ constants C  Cμ ; Cϵ2 ; Cϵ1  and computed a A. Analytical Models of JIC Interactions
joint probability density function (JPDF) for them by calibrating to
There have been many analytical modeling studies of JIC
data from one of the experiments described in Refs. [3–5]. In that
interactions; a comprehensive review can be found in Ref. [14]. The
study, the k − ϵ RANS model employed a linear eddy viscosity
jet, emanating into the crossflow, bends and is swept downstream,
model. The experiment in question consisted of an M  3.93 jet
rolling into a counterrotating vortex pair (CVP). The bulk of the
exhausting into an M  0.8 crossflow, providing a jet-to-crossflow
studies have targeted modeling the trajectory/penetration of the jet in
momentum ratio of J  10.2. We found the calibrated JPDF to be far
the crossflow, as a function of the downstream distance. Broadwell
more predictive than the nominal value of the parameters
and Breidenthal [15] derived a model that showed that in the far-field,
(Cnom  f0.09; 1.92; 1.44g, taken from literature) for some of the
for example, about 50 jet diameters dj from the exit, the trajectory
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

other experiments described in Refs [3,5]. This suggested that,


scaled as y ∼ x1∕3 , a result that agreed with incompressible
perhaps, inappropriate parameters were the main cause of the large
experiments. They also showed that the circulation Γ of the CVP
numerical inaccuracies. decayed downstream as Γ ∼ x−1∕3 . Karagozian [16] developed a
As a check, in our next paper [11], we investigated the second vortex dynamical model of the CVP and rederived the same result.
potential cause of errors, that is, the simplifications inherent in k − ϵ Her model included “viscous” vortices, that is, regions with vorticity
RANS. We estimated a nonlinear eddy viscosity model (i.e., we distributed in space as a Gaussian and which spread in time due to
inferred the model form and associated parameters) using data from viscous processes. She showed that Reynolds number had a weak
the same experiment referred to above. Because this process involved effect on the trajectory. The model was later extended to compressible
discovering both model form and parameters for the k − ϵ RANS crossflows [17] using inviscid point vortices.
model, we expected that it would be substantially more predictive Hasselbrink and Mungal [14] developed scaling laws for the
than our work with the linear eddy viscosity model [9]; instead, we velocities and trajectory of the jet in JIC interactions. They divided
found that their predictive skills to be quite similar. This seemed to the interaction into three phases. The first, called the potential core,
imply that one of the most important drawback of k − ϵ RANS with extended a few dj downstream of the exit and resembled an
linear eddy viscosity models—their inability to accommodate axisymmetric mixing layer. The second, called the near-field,
anisotropy in Reynolds stresses—was not as large a contributing resembled the far-field of a free jet. The trajectory scales as y ∼ x1∕2 .
factor to the inaccuracies in the simulated mean flows compared with The third, called the far-field (about 50 dj downstream of the exit)
inappropriate k − ϵ constants. Note that this is a weak conclusion— displayed scalings that were reminiscent of a wake. The trajectory
the model form that we learned could have been deficient due to a lack scales as y ∼ x1∕3 and the circulation of the CVP obeys Γ ∼ J1∕3 x−1∕3 .
of experimental data. The modeling efforts described above started from first principles
These two studies seemed to indicate that parametric uncertainties, rather than the k − ϵ model and are thus not couched in terms of
rather than model-form errors, were the primary cause of inaccuracies Cμ ; Cϵ2 ; Cϵ1 . However, in one of our conference papers [13], we
in JIC simulations. However, because both the investigations derived an analytical model for JIC interactions that drew on the
involved calibration (fitting to data), there was a possibility that the k − ϵ model and the near-field/far-field formalism described in
calibrated JPDF could be compensating for model-form errors. While
Ref. [14]. The scalings described above, that is, y ∼ x1∕2 , y ∼ x1∕3
the predictive skill of the JPDF for experiments not used in the
calibration seemed to suggest that the degree of compensation was and Γ ∼ J1∕3 x−1∕3 for the near- and far-fields were recovered, but
not large, it was hardly conclusive; the JPDF was not sharp and the now also involved Cμ , Cϵ2 , and Cϵ1 . This allowed us to compute the
estimate of C had a significant degree of uncertainty. In our current trajectory and circulation of the CVP as a function of the analytical
study, we investigate whether it may be possible to find a new value of estimate of C  Cμ ; Cϵ2 ; Cϵ1   Ca and compare with compress-
C that would be predictive for JIC simulations over a range of ible experiments described in Refs. [3,5]. This was done in our
crossflow Mach numbers M and jet-to-crossflow momentum ratios J. conference paper [13] where we found good agreement with
This implies that the model-form errors are small over an M; J M  0.6, 0.7 and 0.8, J  10.2 experiments. Our model was
range. It also provides an indirect check on our previous (weak) derived under an incompressibility assumption, implying that the
conclusion that failure to model anisotropy in turbulent stresses was a effect of compressibility on the JIC interaction in the experiments
smaller source of prediction errors vis-à-vis the use of inappropriate we will calibrate to is weak.
parameters. We will develop and compare JPDFs for the experiments
in Refs. [3–5]. We do not expect the JPDFs to be identical, because B. Improving k − ϵ Models
the model-form errors will be different in each case. However, any k − ϵ RANS models contain grossly simplified representations of
points of similarity between the JPDFs would provide strong turbulent processes, along with numerical parameters. The numerical
evidence of how Cnom should change in order to be predictive. We parameters are estimated from turbulent boundary layers and free
then develop an analytical model of JIC interactions, and obtain an shear flows [18]; we will refer to them as the “nominal” values Cnom .
estimate for C, which we refer to as Ca . While the analytical model They are not universal, and RANS model parameters have been
involves its own set of simplifications, it does not involve fitting to repeatedly tuned for particular classes of flows [19]. k − ϵ models
experimental data, and is thus free of any conflation and have also been augmented with extra terms, for example, for
compensation of parametric versus model-form uncertainties. We axisymmetric and compressible jets [20,21]. The fitting of k − ϵ
compare the predictive skill of Ca versus the calibrated JPDFs, and models to measurements of velocity profiles of supersonic
decide whether it might serve as an alternative to Cnom . axisymmetric jets revealed model parameters that are quite different
This paper is an extension of two of our conference papers [12,13] from Cnom . In particular, their estimate of Cϵ2 is close to its upper
that contain many of the details, including the derivations. We will bound of 2.1.
make references to these details in the paper, which is structured as Recently, the estimation of RANS parameters has tended to adopt
follows. In Sec. II, we review the literature on studies that seek to a Bayesian approach, computing the parameters’ JPDF from
improve the predictive skill of RANS models by using observational experimental and direct numerical simulation (DNS) data. This
data. We also review previous attempts at modeling JIC interactions decision is partly driven by the realization that the data may not be
Article in Advance / RAY ET AL. 3

uniformly informative on the RANS constants being estimated, or equation (Spalart–Allmaras RANS model) were estimated using full-
they may be affected by the inherent shortcomings of RANS models. field (Bayesian) inversion employing DNS [33] and experimental
The first investigations targeted simple flows, for example, flat-plate [34] data of flows over airfoils. They were related to local flow
boundary layers and wall-bounded flows [22,23]. Bayesian model properties using artificial neural networks. Recently, Edeling et al.
averaging has also been used to “merge” JPDFs of RANS constants developed transport equations for anisotropy perturbation to
obtained from a variety of flow cases (favorable and adverse pressure turbulent stresses and reduced the field inversion to that of estimating
gradients) and a host of k − ω, k − ϵ and Spalart–Allmaras models 2 parameters; they demonstrated their method by predicting flow
[24]. Bayesian calibration of RANS constants has also been explored over a backward facing step and a subsonic jet flow [35]. Thus there
in more complicated (3D) geometries. In Ref. [25], the authors exist methods to enrich RANS equations to reduce model-form
calibrated a k − ϵ model for urban canyon flows, whereas in errors, although they may require the solution of a challenging
Refs. [9,11], we addressed compressible jet-in-crossflow inter- inverse problem.
actions. The method developed in Ref. [9] is the one employed for Ling et al. [36] estimated turbulent stress anisotropy by comparing
Bayesian calibration results presented in this paper. These more DNS and RANS simulations and modeled the anisotropy using a
complicated and computationally expensive RANS simulations had neural network that specifically constrained the modeled tensor to be
to employ statistical emulators with Markov chain Monte Carlo Galilean invariant. The neural net model was integrated into a k − ϵ
(MCMC) sampling to construct the JPDF and also take recourse to RANS simulator and used to predict phenomena that cannot be
informative priors, because RANS constants arbitrarily picked from captured by LEVMs. Ling’s approach does not require one to solve an
within their bounds will not necessarily provide realistic flowfields inverse problem. It has been used to model turbulent wall fluctuations
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

(or even allow the simulation to complete). Bayesian estimation of [37] and learning turbulent diffusivity for film cooling flows [38]. Such
parameters of k − ω − γ models, for transitional hypersonic flat-plate methods could be used to improve RANS once shortcomings due to
boundary layers, have successfully used generalized polynomial the use of inappropriate parameters have been be rectified.
chaos expansions to construct statistical emulators and infer six
parameters from experimental measurements of Stanton number III. Inverse Modeling
[26]. The approach is very similar to the method used in this paper, A. Flow Problem
though we do not use polynomial chaos expansion emulators. The wind-tunnel experiments and measurements that have been
Bayesian estimation of k − ω RANS constants have also been used in this paper are fully described in Refs. [3,5]. The
performed using iterated Ensemble Transform Kalman Filters using computational details of the CFD simulator used in this paper are in
measurements from an experiment of a flow over a backward facing our previous paper [9]; consequently, we only provide a summary
step [27]. The use of an “online” method such as Kalman filters, here. A schematic diagram of the wind-tunnel test section is in Fig. 1
which assimilate data incrementally for Bayesian estimation of (left). A jet of diameter 9.53 mm is introduced, at Mach 3.93, into a
turbulence model parameters, can be helpful when calibration crossflow from an orifice located at the front (left side) of the test
datasets are too large for “batch” methods such as MCMC. DNS section floor. Separate experiments are conducted by varying the
simulations provide such datasets. crossflow Mach number M  f0.6; 0.7; 0.8g. The ratio of the jet
Most RANS models employ a linear eddy viscosity model momentum to that of the crossflow, J, is held fixed at 10.2 in
(LEVM) that cannot reproduce the anisotropy observed in turbulent these experiments. Experiments are also performed by varying J 
flows. Much work on data-driven estimation of corrections/ f10.2; 16.7g, with the crossflow Mach number M fixed at 0.8. PIV
augmentation of RANS models aim to rectify this shortcoming. In measurements of the mean flow velocity are made on two planes—
Ref. [28], the authors estimated spatially variable corrections, the midplane (the plane of symmetry in Fig. 1 left), and the cross-
modeled as a Gaussian random field, to the eddy viscosity coefficient plane, the transverse plane that slices through the CVP as seen in
using DNS data. In addition, Ref. [29] developed a genetic algorithm Fig. 1 right). We also plot a window W where quantifications of
to discover stress-strain rate relationships using DNS datasets of flow vorticity will be done in Sec. V. The cross-plane is situated 321.8 mm
over backward-facing step and tested them by predicting flows over downstream of the jet, and the oppositely signed vortices are clearly
periodic hills. In Ref. [30], the authors developed a parametric model seen in the figure.
for the anisotropy in turbulent stresses; the parameters are functions Midplane mean-flow velocity measurements are made at 5
of space. These have been modeled as Gaussian random fields and streamwise locations, the first being 200 mm downstream of the jet,
inferred from DNS data [31]. These corrections for anisotropy were and the rest at 50 mm intervals proceeding downstream. At each of
related to local flow properties using random forests [32]. In other these five locations, velocity measurements are available at 63 points,
work, spatially variable corrections to the eddy viscosity evolution distributed vertically, which we refer to as “probes.” Midplane

0.15

0.1
Y (m)

0.05

0
0.06 0.04 0.02 0 -0.02 -0.04 -0.06
Z (m)
Fig. 1 Left: Schematic of the wind-tunnel test section, which also serves as the computational domain. The orifice where the jet is introduces is also shown,
as are the midplane and the cross-plane. Right: Contour plot of the vorticity field on the cross-plane slicing through the CVP. Positive vorticity is plotted
with red contours. We also plot a window W, where quantifications of vorticity will be performed in Sec. V.
4 Article in Advance / RAY ET AL.

measurements are available for all the experiments. Cross-plane Assume that our prior belief regarding C and σ 2 can be expressed
measurements are available only for the M  0.8 experiments. as the distributions Π1 C and Π2 σ 2 . By Bayes’ theorem, these can
Because they are available for all the experiments, we will use be combined into an expression for PC; σ 2 jye , the posterior density
midplane measurements of streamwise and vertical velocities in our (alternatively, the JPDF):
inverse problem to infer C. We will check the predictive skill of the
calibrated C using measurements on the cross-plane. Note that this PC; σ 2 jye  ∝ Lye jC; σ 2 Π1 CΠ2 σ 2 
necessarily means that we can perform the check only for the subset  
of the experiments where both types of measurements are available. 1 ky − ym Ck22
∝ Np exp − e Π1 CΠ2 σ 2  (2)
Note also that this is different from our previous work [9] in which σ 2σ 2
calibration was performed using cross-plane measurements obtained
solely from the M  0.8; J  10.2 experiment. The actual calibration variables are the velocity deficit in the
The calibration of C employs a k − ϵ RANS model, with a streamwise direction udef and the normalized vertical velocity on the
compressibility correction; details are in Refs. [9,39]. As verified in midplane vnorm . We formally define them as
Ref. [9], the most important k − ϵ constants for the JIC interaction are
C  Cμ ; Cϵ2 ; Cϵ1 . They will be the targets of our calibration study. Umax x − u v
Our flow simulator, SIGMA CFD (Sandia Implicit Generalized udef  and vnorm 
U∞ U∞
Multi-Block Analysis Code for Fluid Dynamics), employs Roe-TVD
fluxes with a minmod limiter to approximate the spatial terms in the where Umax x is the maximum streamwise velocity at an x-location
k − ϵ RANS equation. Integration in time is performed using a first-
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

and U∞ is the freestream velocity. Because of the mass added by the


order point-implicit scheme. The computations start with first-order
jet, the maximum velocity observed in the test section varies in the
spatial discretization, proceed for 5000 time steps, and are then
streamwise direction. udef and vnorm are the elements of ye . Define um
switched to a second-order method for another 25,000 time steps.
to be the predicted streamwise velocity deficit udef and vm be the
Convergence to a steady state is carried out using local time-stepping,
as the CFL is gradually ramped up. We use a multiblock mesh with normalized vertical velocity vnorm for all the probes. Define ue and ve
about 10 million cells; mesh convergence studies are in Ref. [8]. to be their experimental counterparts over all the probes. Therefore,
Initial and boundary conditions are described in detail in one of our Eq. (2) can be written as
previous papers [9].  
1 kue − um C∕K u k22
PC; σ 2 jue ; ve  ∝exp −
σ Np 2σ 2
B. Formulation of the Bayesian Inverse Problem   (3)
In this section, we formulate a Bayesian inverse problem to infer a kve − vm C∕Kv k2
2
× exp − Π1 CΠ2 σ 2 
JPDF for C  Cμ ; Cϵ2 ; Cϵ1  conditional on streamwise and vertical 2σ 2
mean flow velocities measured on the midplane. Because it is the
plane of symmetry, there is no spanwise velocity. Henceforth any where K u  maxue  and K v  maxve . Normalization using Ku
references to a calibrated set of k − ϵ constants C will imply the and K v ensures that the contributions from udef and vnorm are equally
JPDF, rather than a deterministic or point estimate of C. Computing C weighted in the expression for the posterior density. Solving Eq. (3)
as a JPDF captures the uncertainty in the estimate due to limited requires us to specify Π1 C and Π2 σ 2 . As in Ref. [9], we represent
measurements, measurement noise, and the shortcomings of the our prior belief of σ 2 in terms of its reciprocal, that is, Π2 σ −2  which
RANS model that prevent it from reproducing experimental we model with a Gamma distribution, that is, σ −2 ∼ Γk; θ, where
measurements perfectly. The development closely parallels the k  1, θ  1. Such a prior distribution is called an inverse Gamma
formulation of the Bayesian inverse problem in Ref. [9], which used prior and has the advantage of being practically noninformative for
measurements on the cross-plane. Cross-plane measurements are σ −2 > 5. This ensures that our prior beliefs do not influence the
available for only a subset of the experiments described in Refs. [3,5], inference of σ 2 excessively, and the JPDF should reflect the
necessitating the development of this alternative inverse problem information on σ 2 contained in the experimental measurements.
formulation. The four-dimensional distribution PC; σ 2 jue ; ve  in Eq. (3) is
Consider an experimental dataset consisting of a vector ye, of complicated, and is realized by drawing samples of C; σ 2 . As in
length N p , containing measurements at N p probes. Assume that a Ref. [9], the sampling is performed using a particular MCMC
simulation seeded with the k − ϵ constants C results in numerical algorithm called Delayed Rejection Adaptive Metropolis [40]. The
predictions ym C at the same probe locations. We relate the
conjugate inverse Gamma prior for σ 2 allows us to sample for σ −2 via
measurements to the model predictions; thus: ye  ym C  ϵ,
a Gibbs sampler. The MCMC chain is run till it converges to a
where ϵ  fϵi g, i  1; : : : ; N p is the model–data mismatch. It is a
stationary distribution, as assessed using the Raftery-Lewis method
composite of measurement and model-form errors. We define model-
form errors as the discrepancy between noise-free observations and [41]. The software for computing PC; σ 2 jue ; ve  is written in R [42].
model predictions using a value of C that is optimal across all JIC We use the Delayed Rejection Adaptive Metropolis algorithm
interactions. We make the inverse modeling assumption that the implemented in the R package FME [43]. We also use the R package
error at each probe is independent of the others, and can be modeled mcgibbsit [44] for its implementation of the Raftery-Lewis
as independently and identically distributed Gaussians, that is, convergence metric. The samples of C; σ 2  are used to construct the
ϵi ∼ N 0; σ 2 . This statistical representation for the model–data JPDF PC; σ 2 jue ; ve  empirically via kernel density estimation [45].
mismatch introduces an extra parameter σ, which will also be The MCMC method takes between 25,000 and 50,000 samples of
estimated as a part of the calibration; its provides a crude measure C; σ 2  to reach convergence. Because each sample requires one to
of disagreement between model predictions and experimental run a 3D JIC RANS simulation to produce um C; vm C, the
observations. If the measurement errors in the experiment are small, σ MCMC method is impractical as described above. Instead, we
is a measure of the model-form error in k − ϵ RANS when simulating replace the RANS simulator with a statistical emulator. The emulator
the JIC experiment. is a computationally inexpensive proxy of SIGMA CFD that can
We now develop the expression for the JPDF of C; σ 2  conditional provide accurate approximations of um C; vm C within the
on the experimental data ye , that is, PC; σ 2 jye . Given our Gaussian parameter space C defined by:
model for the model–data mismatch ϵi ∼ N 0; σ 2 , the likelihood
Lye jC of observing ye when the CFD simulator is seeded with C is 0.06 ≤ Cμ ≤ 0.12;
  1.7 ≤ Cϵ2 ≤ 2.1;
1 ky − ym Ck22
Lye jC; σ 2  ∝ exp − e (1)
σ Np 2σ 2
1.2 ≤ Cϵ1 ≤ 1.7 (4)
Article in Advance / RAY ET AL. 5

C. Priors and Statistical Emulators fits are constructed for udef and vnorm , and probes are also treated
An informative prior Π1 C: The bounds on Cμ , Cϵ2 , and Cϵ1 individually. At each probe, we postulate a cubic polynomial in
defined in Eq. (4) could be used to propose a prior distribution Π1 C, Cμ ; Cϵ2 ; Cϵ1  as a model for udef (or vnorm ). We use the values of C that
for example, as a uniform distribution on C. However, an arbitrary lie in R as the data in a least-squares fit to estimate the polynomial’s
point in C may not lead to a physically realistic combination coefficients. The polynomial is simplified using Akaike Information
Cμ ; Cϵ2 ; Cϵ1 —the RANS simulation may not run, fail to converge Criterion. We find that many of the cubic terms are removed, and in
to a steady state, or predict a flow field that does not resemble case of some probes, the polynomial model reduces to a quadratic.
transonic, high-Reynolds-number flows. Thus, in order to construct Before using the polynomial model as a replacement for SIGMA CFD
Π1 C, we will first need to isolate a physically realistic region R ⊂ C in the inverse problem, we ensure that it is an accurate proxy. We
in the parameter space. In doing so, we will follow the method compute the approximate error of the polynomial emulator using
developed in our previous paper on RANS calibration [9], and so we repeated random subsampling, a form of cross-validation; only those
summarize the process here. emulators that achieve an approximation error less than 15% are
The prior Π1 C is constructed during the process of devising a retained. We see that this process often removes about 2/3rd of
statistical emulator for SIGMA CFD. We distribute 2744 143  polynomial proxies, and the probes with accurate emulators follow the
samples of C inside C using a quasi-random, space-filling Halton trajectory of the jet. This is not unexpected—the impact of varying C is
sequence and seed RANS JIC simulations with them. Of these a few felt in regions with substantial velocity gradients and turbulence. The
simply fail to converge to a steady state (due to nonphysical run-to-run variation in the regions of the flowfield away from the jet is
Cμ ; Cϵ2 ; Cϵ1  combinations), whereas others run to completion but small and mostly due to numerical noise, which cannot be modeled
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

may provide unsteady or unrealistic flowfields. We compute the root with Cμ ; Cϵ2 ; Cϵ1 . An illustration of such “model-able” probes
mean square (RMS) discrepancy between ue ; ve  and um ; vm  tracking the jet trajectory can be found in Ref. [12], Fig. 2. Note that as
produced by each of the RANS simulations, and retain the top 25% of in Ref. [9], no emulators are built for probes inside the boundary layer;
the run, that is, the ones closest to the experimental measurements. the calibration is aimed at the accurate simulation of the jet, the CVP
The values of C so chosen are deemed to constitute R, the physically and their trajectory.
realistic subset of the parameter space C. As in Ref. [9], we define The procedure outlined in Secs. III.B and III.C is repeated for all
 four experiments considered in this study, resulting in four JPDFs
1 if C ∈ R PC; σ 2 jue ; ve . However, to be useful, we have to ensure that the
Π1 C  (5)
0 otherwise information content in ue ; ve  on Cμ ; Cϵ2 ; Cϵ1  is substantial,
preferably similar to that of the cross-plane measurements, as
For the particular case of the M  0.8; J  10.2 experiment, of assessed in Ref. [9]. We check this in two ways. First, we compare the
the 2744 runs, 2628 completed. Of these the top 25%, that is, 657, were JPDF PC; σ 2 jue ; ve  for the M  0.8; J  10.2 experiment with
used to define R; they are plotted in Fig. 2. We note that they constitute the one inferred from cross-plane measurements in Ref. [9]. This is
a contiguous region in C. The discrete values of Π1 C, as initialized shown in Fig. 3. There, the posterior density obtained using midplane
using Eq. (5), constitute the training data (TD) to construct a binary measurements is plotted using a solid line, whereas the one from
support vector machine classifier (SVMC). For the particular case Ref. [9], inferred using cross-plane measurements, is plotted using a
of the M  0.8; J  10.2 experiment, the TD consisted of 2628 dashed line. Prior densities, which are the projections of R onto the
records, of which 657 fell in one class. The process for constructing axes, are plotted with ∘ and Δ for the midplane and cross-plane cases,
the SVMC, and the R package used to do so, are described in respectively. We see that the prior densities are very close, indicating
Ref. [9]. We ensure that the SVMC has a misclassification rate of less that the physically realistic parameter space R identified in this
than 10% and thus can define R accurately inside C. We use this paper using udef and vnorm may be very similar to the one identified in
SVMC as an implementation of the prior Π1 C in Eq. (3), integrate it Ref. [9] using cross-plane measurements. Also the prior distribution
into the MCMC software, and use it to sample posterior density is sufficiently wide to comfortably contain the posterior density.
PC; σ 2 jue ; ve . We take this as a sign that the procedure for identifying and enforcing
Statistical emulator for SIGMA CFD: We use polynomial curve-fits R, using the 25% threshold described above, is not restrictive.
to map the inputs into SIGMA CFD, that is, Cμ ; Cϵ2 ; Cϵ1 , to the Further, the posterior densities are almost identical, indicating that
outputs udef and vnorm . These curve-fits are then used as computa- the information on Cμ ; Cϵ2 ; Cϵ1  contained in the midplane
tionally inexpensive proxies for our RANS simulator. Separate curve- measurements may be similar to the information contained in the
2.1

1.55

1.55
1.50

1.50
2.0

1.45

1.45
1.40

1.40
C2

C1

C1
1.9

1.35

1.35
1.30

1.30
1.8

1.25

1.25
1.20

1.20
1.7

0.06 0.08 0.10 0.12 0.06 0.08 0.10 0.12 1.7 1.8 1.9 2.0 2.1
Cμ Cμ C2
Fig. 2 C points in C that constitute R, projected down to the Cμ ; Cϵ2 , Cμ ; Cϵ1 , and Cϵ2 ; Cϵ1  planes.
6 Article in Advance / RAY ET AL.

40
Posterior; mid−plane

10 12
Prior; mid−plane

30
Posterior; cross−plane

30
Density Prior; cross−plane

Density

Density
8
20

20

6
4
10

10

2
0

0
0.06 0.07 0.08 0.09 0.10 0.11 0.12 1.7 1.8 1.9 2.0 2.1 1.2 1.3 1.4 1.5 1.6 1.7
Cμ C2 C1
Fig. 3 Prior and posterior densities obtained from midplane and cross-plane measurements. Posterior densities obtained from midplane and cross-plane
measurements are plotted using solid and dashed lines, respectively. The corresponding priors are plotted with ∘ and Δ. The vertical lines are the
“nominal” values of the parameters.

cross-plane measurements. The vertical lines in the figures denote the conference paper [12] and the two ensembles match very closely.
“nominal” values of Cμ , Cϵ2 , and Cϵ1 , and are quite different from the
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

This is not surprising, given the agreement between the two JPDFs.
maximum a posterior (MAP) values (the peak of the PDF). This gives us some confidence that the midplane and cross-plane
The close agreement between the two posterior densities in Fig. 3 measurements contain similar levels of information on Cμ ; Cϵ2 ; Cϵ1 .
implies that they are equally predictive. This can be checked using a This is important because two of the four experiments considered in
“pushed-forward posterior” test and is documented in Ref. [12]. We this paper, namely, the M  0.6; J  10.2 and M  0.7; J 
seed our RANS simulator (SIGMA CFD) with 100 samples of C 10.2 experiments, do not have cross-plane measurements and we have
drawn from PC; σ 2 jue ; ve  and develop an ensemble of velocity no independent way of checking whether the experiments’ JPDF
predictions on the cross- and midplanes. These are compared with can reproduce flowfield measurements that were not used in the
flowfields developed using a similar “pushed-forward posterior” test calibration.
with samples drawn from the JPDF inferred using cross-plane The bulk of the computational cost is incurred in generating the
measurements. The comparison is available in Figs. 4 and 5 in our TD. Each SIGMA CFD simulation of a JIC interaction requires

M = 0.6, J = 10.2
250

M = 0.7, J = 10.2
40

M = 0.8, J = 10.2
M = 0.8, J = 16.7
200

30
Density
Density
150

20
100

10
50

0
0

0.06 0.07 0.08 0.09 0.10 0.11 0.12 1.7 1.8 1.9 2.0 2.1
Cμ C2
60
150

50
100

40
Density

Density
30
20
50

10
0

1.2 1.3 1.4


1.5 1.6 1.7 0.10 0.15 0.20 0.25 0.30
C1 σ
Fig. 4 Marginalized JPDFs of Cμ ; Cϵ2 ; Cϵ1 ; σ obtained from the experimental data in Refs. [3]. The dashed vertical line is the nominal value of the k − ϵ
constant and the solid vertical line is the analytically derived value [Eq. (31)].
Article in Advance / RAY ET AL. 7

60
04
Density
20
0

0.06 0.10

2.10

80
60
Density
2.06

40
C2

20
2.02
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

0.06 0.08 0.10 0.12 2.02 2.06 2.10


Cμ C2
1.45

1.45

15
10
Density
1.35
1.35

C1
C1

5
1.25

1.25

0.06 0.08 0.10 0.12 2.02 2.06 2.10 1.20 1.30 1.40 1.50
Cμ C2 C1
0.12 0.14 0.16 0.18

0.12 0.14 0.16 0.18


0.12 0.14 0.16 0.18

10 20 30 40
Density
σ

0.06 0.08 0.10 0.12 2.02 2.06 2.10 0.16 1.25 1.35 1.45 0.12
Cμ C2 σ C1
Fig. 5 The JPDF obtained from the M  0.7; J  10.2 experimental dataset, marginalized to two dimensions. The correlated nature of Cϵ1 ; Cμ 
is clear.

about 12 h on 1024 cores of a PowerPC A2 processor; 2744 runs (streamwise) location is the y (vertical) location where the vertical
are required to generate the TD for a given experiment. We generate normalized velocity vnorm reaches a maximum. We represent y 
TDs for four separate experiments corresponding to M  0.6; Cxn as the trajectory and fit this model to the RANS data. In Fig. 6, we
J  10.2, M  0.7; J  10.2, M  0.8; J  10.2, and plot C  y∕x1∕3 using ∘ and C  y∕x1∕2 using ×. We clearly see the
M  0.8; J  16.7. The “pushed-forward posterior” tests require two scalings fitting the trajectory. For x∕dj > 13 (the vertical line), C
additional runs. Simulations were performed on the Sequioa assumes a constant value of around 2.9; further, for x∕dj < 13,
supercomputer (https://asc.llnl.gov/computing_resources/sequoia/) C varies between 2 and 2.9. On the other hand, when we plot
at Lawrence Livermore National Laboratory. C  y∕x1∕2 , C assumes a value of around 1.8 for x∕dj < 13, and then
slowly decays. Thus the separation of near-field and far-field
IV. Analytical Model behaviors, described in Ref. [14], is replicated here. This gives us
confidence that an analytical model that draws from a k − ϵ model
Hasselbrink and Mungal [14] described, in a JIC interaction, the
could accurately represent the data in Ref. [3,5].
existence of a near-field, where the trajectory of the jet scaled as
y ∼ x1∕2 , and a far-field where the scaling proceeded as y ∼ x∕13 . We
will use their formalism to develop our analytical model, starting A. Far-Field Model
from the k − ϵ RANS equations, so that we may obtain analytical We develop a self-similar model for the far-field first. It follows the
estimates for Cμ, Cϵ2 , and Cϵ1 . We first check whether k − ϵ RANS development of a similar model for wakes [10], in keeping with its
equations show these scaling for compressible JIC interactions. In similarity with the far-field behavior of JIC [14]. We assume that the
Figure 6 we plot the scaled trajectories for a J  10.2 jet interacting far-field interactions are only weakly compressible, and we will
with an M  0.6 crossflow, simulated using the nominal values of the develop the model under incompressibility and axisymmetric
RANS constants Cnom . The simulation domain and settings are the assumptions where the x − r plane contains the jet. We also assume
same as the one described in Sec. III.A. The location of the jet at any x that the jet, which was introduced vertically, has curved into the
8 Article in Advance / RAY ET AL.

Trajectory, scaled u^ s  Axn−1 and l  Bxn (10)


3
2.8 The streamwise momentum flux integrated across the jet at any x
2.6
location is
n = 1/3
n = 1/2
Z ∞
Z ∞
2.4
_
2π m ur dr  2π ml
_ 2 U∞  uξ
^ dξ (11)
2.2 0 0
n
y /x

2 where m _  ρ∞ U∞ is the mass flux and ρ∞ is the freestream


crossflow density. This momentum consists of the momentum due to
1.8
the crossflow mU _ ∞ and that due to the jet. Equating the jet’s
1.6 contribution to the momentum to the jet’s momentum at the jet exit at
the bottom of the test section, we get
1.4
Z
π 2 ∞
1.2 d ρ V 2  2πl2 u^ s ρ∞ U∞ ξf dξ (12)
4 j j j 0
1
0 5 10 15 20 25 30 35 40 45
Since the left hand side is a constant, l2 u^ s too should be a constant.
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

x/dj
From Eq. (10) we get
Fig. 6 Trajectory of an M  0.6; J  10.2 JIC interaction, simulated
using k − ϵ RANS equations, scaled as y  Cxn . We plot C  y∕x1∕3
using ∘ and C  y∕x1∕2 using ×. There are two significantly different 3n − 1  0 or n  1∕3 (13)
regimes for the trajectory, separated by the vertical line at x∕dj  13.
Combining Eqs. (13) and (9), we get

crossflow to an extent that the vertical velocity in the jet is small, that u^ s  A U∞ d2∕3 −2∕3 and l  B d2∕3 1∕3
j x j x (14)
is, v∕U∞ ≪ 1. We also assume that the mass flow and momentum of
the jet (in magnitude) is the same as when it was introduced into the
crossflow. The center of the jet has a streamwise velocity deficit that where A and B are nondimensional. We substitute Eq. (14) into
changes downstream, and we will model the velocity deficit profile at Eq. (8) to get
a given x location, in a manner similar to wakes. The details of the  
derivation are in Ref. [13]. B D 0 1 ∂ 0g
2
− 2f  ξf   ξf (15)
Let u  U∞  u, ^ ∞ ≪ 1. We also set
^ u∕U 3A C2 Cμ ξ ∂ξ h

u^  u^ s xfξ; k  ks xgξ and ϵ  ϵs xhξ (6) Following the modeling of wakes [46], we set

where l  lx is a turbulence length scale and ξ  r∕l. Under the D g2


assumption of incompressibility and negligible vertical velocity, the  ReT  14.1 and 1 (16)
C2 C μ h
momentum equation can be linearized as
  where ReT is the turbulence Reynolds number and is set to 14.1 per
∂u 1 ∂ k2 ∂u
U∞  rCμ (7) Ref. [47]. In Eq. (16), set
∂x r ∂r ϵ ∂r
B D B
where we have used a linear eddy viscosity model for the turbulent −  α  1; to get −  ReT − 1 (17)
3A C2 Cμ 3A
kinematic viscosity νT  Cμ k2 ∕ϵ. We write

∂u ∂u^ l0 From Eq. (15) we get


  u^ s0 f − u^ s ξf 0
∂x ∂x l  2
1 ∂ ξ
2f  ξf 0  ξf 0 ; implying f  exp − (18)
and substitute into Eq. (7) to get ξ ∂ξ 2
   
l0 k2 u^ 1 ∂ g2 Using Eq. (12), we also get
U∞ u^ s0 − u^ s ξf 0  Cμ s 2s ξf 0 (8)
l ϵs l ξ ∂ξ h
J
 A B 2 (19)
Here prime ( 0) denotes a first derivative with respect to the 8
independent variable that is x in case of l 0 and ξ in case of f 0 . We
define ks  Cu^ 2s and ϵs  Du^ 3s ∕l, which we substitute into Eq. (8), where J is the jet-to-crossflow momentum ratio. Using Eqs. (19) and
and divide both sides by u^ 2s ∕l, to get (17) to compute A and B and substituting into Eq. (14), we get
     
u^ s0 f − u^ s ξf 0 l 0 ∕l Cμ 1 ∂ 2 C2 u^ s 1 ReT 2∕3 1∕3 d 2∕3
 ξf 0g  J
u^ 2s ∕l U∞ ξ ∂ξ h D U∞ 2 3 x
   
l 1 3 1∕3 1∕3 x 1∕3
The right-hand side of the equation is solely a function of ξ, which  J (20)
dj 2 ReT d
implies

l ∂u^ s 1 ∂l B. k − ϵ Parameter Estimates


 K1 and  K2 (9)
u^ 2s ∂x u^ s ∂x We use the far-field profiles developed in Sec. IV.A to derive
expressions for Cϵ2, Cμ , and Cϵ1 . We linearize the evolution equations
Equation (9) holds if for k and ϵ as we had done for Eq. (7) to get
Article in Advance / RAY ET AL. 9

   
∂k 1 ∂ Cμ k2 ∂k k2 ∂u 2 D
U∞  r  Cμ −ϵ Cϵ2 ReT g  4
∂x r ∂r σ k ϵ ∂r ϵ ∂r C

for the k− equation and, for the ϵ− equation: Combining with Eq. (22), we get

   2 Cϵ2  2 (28)
∂ϵ 1 Cμ k2 ∂ ∂ ϵ2
U∞  r ϵ  Cϵ1 Cμ k u − Cϵ2
∂x r σ ϵ ϵ ∂r ∂r h and the profile for ϵ is given by

Introducing the expressions in Eq. (6), with ks  Cu^ 2s and 1 ∂ C


ϵs  Du^ 3s ∕l, with u^ s and l replaced by the expressions in Eq. (20), ξh 0  ξh 0   ϵ1 f 0 2  0
ξ ∂ξ C
we get,
Comparing with Eq. (23), we see that h  Cϵ1 g. Recall from
1 ∂ξg 0 1 0 2 D Eq. (16) that g2 ∕h  1, that is, h  g2 ≈ gg.
 Therefore,
α2g  ξg 0    f  − ReT h  0
ξ ∂ξ C C
1 ∂ D h2 h 2 ln 2 − 1
α4h  ξh 0  
C
ξh 0  ϵ1 f 0 2 − Cϵ2 ReT 0 (21) g ≈  Cϵ1  (29)
ξ ∂ξ C C g g 4C
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

where we use Eq. (25) for the numerical value of g.  To complete the
Recall that in Eq. (16), we had stated that g2 ∕h  1. We
estimates of Cμ ; Cϵ2 ; Cϵ1 , we need C. We use the Bradshaw
approximate this as h  g2 ≈ gg, where g is the mean value of g over
the k-profile. We substitute in Eq. (21) (the k− equation) and also set assumption for turbulence shear stress [46], that is, u 0 v 0  2∕3×
α  1, to get 0.45 × ks  0.3ks . Also u 0 v 0 ≈ u^ 2s f 0 ∕ReT where f 0  0.5fmax
0

f 0 ∞ ≈ 0.31 (a simple mean over the entire u-profile). This gives:
1 ∂ 1 D
2g  ξg 0  ξg 0  f 0 2 − ReT gg
 0 0.31u^ 2s
ξ ∂ξ C C u 0v 0   0.3ks
ReT
If we demand that
Consequently
D
 T 2
gRe (22) ks 0.31Re−1
C C  T
 0.072
2 0.3
u^ s
then the equation for k simplifies to
and
01 ∂ 1
ξg  ξg 0  f 0 2  0 (23)
ξ ∂ξ C 2 ln 2 − 1
Cϵ1   1.34 (30)
4C
leading to a solution for g
Collating Eqs. (26), (30), and (28), we get
  2 
1 ξ
g 2Ei − 2Eiξ2  − exp−ξ2  (24) Cμ  0.1; Cϵ2  2.0 and Cϵ1  1.34 (31)
2C 2

where These are the analytical values of C, that is, Ca . They depend on the
modeling assumptions in Eqs. (16), (17), (22), and (27), whose
Z ∞ validity is established by the predictive skill of Ca . Equation (16) is
exp−u
Eix  du borrowed from the modeling of wakes; as stated in Ref. [14], scaling
x u laws show that they are similar to the far-field in JIC interactions.
are exponential integrals. This allows g 0 ξ  0  0 and g∞  0
which satisfy the boundary conditions for the k-profile. Also we V. Results
approximate g as
We perform Bayesian calibration, following the method described
1 1 in Secs. III.B and III.C, to construct four separate JPDFs using data
g  g0  g∞  2 ln 2 − 1 (25) from experiments described in Refs. [3,5]. The JPDFs are
2 4C marginalized and plotted in Fig. 4. The experiments M  0.6;
J  10.2, M  0.7; J  10.2, M  0.8; J  10.2, and M 
a simple arithmetic mean over the k-profile. Using Eqs. (16) and (22),
0.8; J  16.7 are plotted using solid, dashed, dotted, and long-
and substituting the numerical value of g from Eq. (25), we get,
dashed lines. The nominal values of Cμ ; Cϵ2 ; Cϵ1  are plotted with a
dashed vertical line and the analytical value Ca using a solid line. It is
8 D 8
Cμ  Re−2  0.1 and  Re−1 (26) clear that the calibrated Cϵ2 and analytical values of Cϵ2 are far larger
2 ln 2 − 1 T C2 2 ln 2 − 1 T than the nominal value, and in line with the value obtained for
axisymmetric jets [18]. Further, one could argue that the upper bound
Doing the same for the equation for ϵ in Eq. (21) we get on Cϵ2 could be increased to 2.5 (instead of 2.1), as used in Ref. [25].
Note that the tendency of the posterior densities to peak near the
1 ∂ C D h2
4h  ξh 0  ξg 0  ϵ1 f 0 2 − Cϵ2 ReT 0 extremities of the parameter ranges is not an artifact of how we
ξ ∂ξ C C g construct Π1 C but rather that the parameter ranges in Eq. (4), taken
from literature, may be too narrow. The PDFs for Cμ do not have the
We demand that same consistency as the PDFs for Cϵ2, though most of the peaks are at
values larger than the nominal value. This conclusion is also
Cϵ2 ReT D∕Ch∕g  4 (27) supported by the analytical value, which is larger than the nominal
one of 0.09. The PDFs for Cϵ1 are all close to the nominal value of
Further, approximating h∕g ≈ gg∕g,
 we get 1.44 with the exception of the M  0.6; J  10.2 case. The plots of
10 Article in Advance / RAY ET AL.

the model–data mismatch σ, which is in terms of the normalized locations x∕dj  21; 31.5 and 42. Results are presented for the
streamwise velocity deficits and vertical velocities, show that the M  0.8; J  10.2 test case. The dotted line denotes predictions
M  0.6; J  10.2 calibration incurs the largest error; this will be using Cnom , the dashed line stands for Ca, ∘ denote experimental
investigated in detail later. The values of σ imply that a “pushed- measurements, and the solid line is the ensemble mean prediction
forward posterior” run seeded by values of Cμ ; Cϵ2 ; Cϵ1  sampled produced by the pushed-forward posterior. The vertical axis
from the JPDFs plotted in Fig. 4 will not bracket the experimental measures the normalized height y∕dj above the test-section floor. We
measurements; instead they will need a model–data mismatch see that the both udef and v∕U∞ reach a maximum in the middle,
ϵ ∼ N 0; σ 2 , where σ is sampled from the PDF in Fig. 4. This aspect coinciding with the center of the jet. Both decay to zero near the top of
of the calibration for Cμ ; Cϵ2 ; Cϵ1  was observed and explained in the test section. The vertical velocity is zero at the floor of the test
our previous paper where we developed the calibration method [9]; it section, whereas the streamwise velocity deficit decreases to near-
is due to a combination of model-form error (the main component) zero as we depart the center of the jet and then increases as we
and the approximation error of the statistical emulator for SIGMA progress into the boundary layer on the floor of the test section.
CFD. The M  0.7; J  10.2 experiment yields bimodal distribu- Comparing the experimental measurements with the predictions
tions for Cμ and Cϵ1 . In Fig. 5, we plot the JPDF, marginalized to two using Cnom and the ensemble mean, we find the JPDF improves the
dimensions. We clearly see that Cμ and Cϵ1 are correlated, achieving predictive skill of the simulations quite remarkably. What is
similar accuracies by compensating for each other. surprising and reassuring is the degree of agreement achieved by the
Next, we check the predictive skill of the JPDFs plotted in Fig. 4. We flowfield predicted using Ca ; further, the improvement holds across
do so using “pushed-forward posteriors” using 100 samples drawn both the streamwise and vertical velocity fields as measured on the
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

from the JPDF (henceforth, the “ensemble”), as described in Sec. III.C. midplane at all three downstream locations. We do not have
The variability in the prediction of midplane and cross-plane quantities experimental measurements any closer to the jet exit to check the
is not much (below 10%) as shown in our previous paper [9], and predictive skill of the JPDFs in the near-field.
consequently we use the ensemble mean for visualizing midplane The good agreement observed in Fig. 7 is expected because
predictions. We also identify the C sample that produces predictions midplane velocities were the calibration variables. We check the
that match the measurements best; we call this Cμ ; Cϵ2 ; Cϵ1  ability of the JPDF for the M  0.8; J  10.2 test case to
combination Copt . Because the variability of midplane predictions due reproduce the vorticity field on the cross-plane; cross-plane
to uncertainty in the inferred Cμ ; Cϵ2 ; Cϵ1  is small, the ensemble measurements are not included in the calibration. In Fig. 8 (left) we
mean prediction is quite close to the one produced by Copt. plot contours of the vorticity field inside W produced by Cnom using
In Fig. 7 we plot the streamwise velocity deficit udef (top) solid lines. In the middle subfigure, we plot the vorticity field
and normalized vertical velocity v∕U∞ (bottom) at 3 streamwise predicted using Ca . In the right subfigure, we plot the vorticity field

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0

14 14 14

12 12 12

10 10 10
y/dj

y/dj

y/dj

8 8 8

6 Experiment 6 6
Ensemble mean
4 Cnom 4 4
Ca
2 2 2

0 0.1 0.2 0.3 0 0.1 0.2 0.3 0 0.1 0.2 0.3


(Umax - u)/U (Umax - u)/U (Umax - u)/U

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0

14 14 14

12 12 12

10 10 10
y/dj

y/dj

y/dj

8 8 8

6 6 6
Experiment
4 4 4 Ensemble mean
Cnom
2 2 2 Ca

0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0 0.1 0.2 0.3
v/U v/U v/U
Fig. 7 Streamwise velocity deficit (above) and normalized vertical velocity (below) computed using Cnom and Ca compared with the ensemble mean from
the “pushed-forward posterior” test and experimental measurements. Results are plotted for the M  0.8; J  10.2 test case, at three streamwise
locations x∕dj  21; 31.5, and 42.0). We see that both Ca and Copt provide better predictions than the nominal parameters Cnom .
Article in Advance / RAY ET AL. 11

Vorticity (Nominal coefficients) Vorticity (Analytical coefficients) Vorticity (best case)

0.12 0.12 0.12

0.1 0.1 0.1

z [m]

z [m]
z [m]

0.08 0.08 0.08

0.06 0.06 0.06

0.04 0.04 0.04

0.02 0.02 0.02


0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

y [m] y [m] y [m]


Fig. 8 Vorticity field on the cross-plane, as predicted using Cnom (left), Ca (middle), and Copt (right) for the M  0.8; J  10.2 test case. The domain
shown here is the window W. Numerical predictions are plotted using a solid line (–). Overlaid are experimental measurements of vorticity using a dash-
dotted (– · –) line. Again, Copt and Ca improve predictions vis-á-vis the experimental measurements.

produced using Copt . Overlaid on all three subfigures is the VI. Discussion
experimentally observed vorticity field plotted with the dash-dotted The midplane results provided in Figs. 7, 9, and 11 show that the
lines. It is clear, again, that there is a large improvement in the calibrated predictions (i.e., ensemble mean of the “pushed-forward
agreement between experimental and modeled vorticity field as we posterior” predictions, seeded using the JPDFs in Fig. 4) invariably
replace Cnom with Copt . Further, a similar degree of improvement is have a better agreement with experimental measurements than
also observed when we use Ca , in Fig. 8 (middle). In addition, the predictions using the nominal value of k − ϵ constants Cnom . This is
predictions using Ca and Copt are very similar. perhaps not very surprising, given that midplane measurements are
We next check the predictive accuracy of the JPDFs for the the calibration variables. What is surprising is the degree to which
M  0.8; J  16.7 test case, using midplane udef and vnorm as well predictions using the analytical values of C, that is, Ca , follow the
as the cross-plane vorticity field as the comparison metrics. The plots calibrated ensemble means in all three test cases analyzed above. As
for the midplane quantities are in Fig. 9. We see that the larger J mentioned before, the JPDFs almost certainly compensate for some
causes problems for simulations when Cnom is used and the RANS model-form errors; however, given the close agreement with
predictions of udef and vnorm get progressively worse as we go Ca predictions, they are probably not excessive. Ca , being derived
downstream. In the Appendix, we check if predictions using C analytically, with no fitting to the datasets used in this paper, is free of
sampled from the prior bracket observations. We see that they do, any conflation between parametric and model-form uncertainties.
except for about 10 probes at the most downstream location at Further, had the compensation of model-form errors by the JPDFs
x∕dj  42.0. The performance of the calibration is almost certainly been large, the residual data–model mismatch σ would have been
affected by this shortcoming of the prior. Predictions using Ca and the small. As the PDFs for σ in Fig. 4 show, this is not the case.
ensemble mean are far better than the predictions with Cnom . Further, The ability of the JPDFs to reproduce the measured flowfield
the predictions with Ca agree closely with the ensemble mean of the on the cross-plane is somewhat mixed. We checked this by identi-
“pushed-forward posterior” simulations, though the agreement fying an “optimal” C  Copt from the “pushed-forward posterior”
worsens with increasing x∕dj . The numerical simulations under- simulations, and comparing its cross-plane vorticity predictions
predict both udef and vnorm , and consequently a weaker CVP. In against experimental data. Note that a different Copt is obtained for
Fig. 10, we plot the vorticity field inside W as predicted using Cnom each test case; these are tabulated in Table 1. Calibration invariably
(left subfigure), Ca (middle subfigure), and Copt (right subfigure). provided a better prediction than Cnom , and for the M  0.8; J 
Predictions using the nominal values of C do not even reside inside 10.2 test case, the match with experimental data was quite
W, and by that token, Ca and Copt perform far better. However, as impressive. The agreement was somewhat less impressive in the
Fig. 9 shows, the agreement between modeled and experimental M  0.8; J  16.7 test case (see Fig. 10), which may be partially
velocity fields midplane degrades as we proceed downstream, and due to the overly restrictive prior (see the Appendix for details). In
consequently comparison between the measured and predicted all cases, predictions using Ca closely tracked the calibrated
vorticity fields on the cross-plane (situated at x∕dj  33.76) is not as predictions.
good as the one observed for the M  0.8; J  10.2 case (compare It is clear, then, that the JPDFs could be improved. In Ref. [9],
Fig. 10 versus Fig. 8; the simulated vortex sits below the experimental where we developed the Bayesian calibration technique, we had
one for J  16.7). This is particularly true for the predictions obtained a JPDF for C using vorticity measurements on the cross-
using Ca . plane using the M  0.8; J  10.2 test case. We had also obtained
In Fig. 11 we plot the midplane predictions using Cnom and the an “optimal” value of C using a genetic algorithm, which we tabulate
ensemble mean from “pushed-forward posterior” simulations below (Table 1) as Cga . The dominant flow feature in the JIC
from the JPDF for the M  0.7; J  10.2 test case. We see that interaction is the CVP and calibrating to its direct measurements
predictions are quite close to the experimental measurements, and are provided us with a robust JPDF. Being able to reproduce the CVP in
a large improvement over predictions using Cnom . This is particularly the correct position and with the correct circulation automatically led
true for the vertical velocity. In Fig. 12, we plot the vorticity field on to a flowfield that agreed with measurements on the midplane. It was
the cross-plane inside W, as predicted using Ca and Copt . We see that also compared, with somewhat less agreement, to measurements
they agree quite well. No experimental measurements were taken on from the M  0.8; J  16.7 and M  0.7; J  10.2 test cases.
the cross-plane for this test case, and so we do not know the degree to However, it was not optimized to those particular test cases, and one
which the calibrated and analytical vorticity fields improved on the could never be sure whether any discrepancies between predictions
one generated by Cnom ; consequently, we have omitted that plot. developed using Cga and Ca were due to compensations of
12 Article in Advance / RAY ET AL.

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


16 16 16

14 14 Experiment 14
Ensemble mean
12 12 Cnom 12
Ca
10 10 10
y/dj

y/dj

y/dj
8 8 8

6 6 6

4 4 4

2 2 2

0 0.1 0.2 0.3 0 0.1 0.2 0.3 0 0.1 0.2 0.3


(Umax - u)/U (Umax - u)/U (Umax - u)/U
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


18 18 18
16 16 16
14 14 14
12 12 12
10 10 10
y/dj

y/dj

y/dj
8 8 8
6 6 6 Experiment
Ensemble mean
4 4 4 Cnom
2 2 2 Ca

0 0.2 0.4 0.6 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
v/U v/U v/U
Fig. 9 Streamwise velocity deficit udef (above) and normalized vertical velocity vnorm (below) computed using Cnom and Ca compared with the ensemble
mean from the “pushed-forward posterior” simulations and experimental measurements. Results are plotted for the M  0.8; J  16.7 test case, at
three streamwise locations x∕dj  21; 31.5, and 42.0). We see that both Ca and Copt provide better predictions than the nominal parameters Cnom . Also,
the agreement between simulations and experimental measurements worsens downstream.

Vorticity (Nominal coefficients) Vorticity (Analytical coefficients) Vorticity (best case)

0.12 0.12 0.12

0.1 0.1 0.1


z [m]

z [m]

z [m]

0.08 0.08 0.08

0.06 0.06 0.06

0.04 0.04 0.04

0.02 0.02 0.02


0 0.01 0.02 0.03 0.04 0.05 0.06 0 0.01 0.02 0.03 0.04 0.05 0.06 0 0.01 0.02 0.03 0.04 0.05 0.06
y [m] y [m] y [m]
Fig. 10 Vorticity on the cross-plane as predicted using Cnom (left), Ca (middle), and Copt (right) for the M  0.8; J  16.7 test case. The domain shown
here is the window W. Numerical predictions are plotted using a solid line (–). Overlaid are experimental measurements of vorticity using a dash-dotted
(– · –) line. Copt and Ca have a better agreement with measurements than the prediction using Cnom , which does not even lie inside W.

model-form errors, the approximations adopted to obtain Ca or the measurements were only available for M  0.8 test cases). The
shortcomings of the JPDF due to limited and noisy measurements. midplane measurements are less informative of the CVP than the
This led to our decision to calibrate to each of the test cases, and the cross-plane ones, and hence our mixed results in reproducing cross-
necessity of using midplane measurements (because cross-plane plane vorticity fields, postcalibration.
Article in Advance / RAY ET AL. 13

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


16 16 16

14 14 Experiment 14
Ensemble mean
12 12 Cnom 12
Ca
10 10 10
y/dj

y/dj

y/dj
8 8 8

6 6 6

4 4 4

2 2 2

0 0.1 0.2 0.3 0 0.1 0.2 0.3 0 0.1 0.2 0.3


(Umax - u)/U (Umax - u)/U (Umax - u)/U
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


18 18 18
16 16 16
14 14 14
12 12 12
10 10 10
y/dj

y/dj

y/dj
8 8 8
6 6 6 Experiment
Ensemble mean
4 4 4 Cnom
2 2 2 Ca

0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
v/U v/U v/U
Fig. 11 Streamwise velocity deficit udef (above) and normalized vertical velocity vnorm (below) computed using Cnom and Ca compared with the ensemble
mean from the “pushed-forward posterior” simulations and experimental measurements. Results are plotted for the M  0.7; J  10.2 test case, at
three streamwise locations x∕dj  21; 31.5, and 42.0). We see that both Ca and Copt provide better predictions than the nominal parameters Cnom .

the ensemble mean of the “pushed-forward posterior” simulations,


without any measure of the variability in them. This is because the
0.12 variability is small and will not help bracket the observations; as
shown in Ref. [9], one needs to include the effect of σ to successfully
do so. This implies that, postcalibration, the parametric uncertainty is
0.1 far smaller than σ, a composite of model-form, measurement, and
surrogate modeling errors, and any further improvement in predictive
skill is dependent on reducing σ, for example, via reducing model-
z [m]

0.08 form errors.


Figure 4 shows two curious features—the estimate of Cϵ1
obtained for the M  0.6; J  10.2 case is quite different from the
0.06
other ones and the data–model mismatch σ is the largest of all the
test cases considered. We investigate this further. In Fig. 13, we plot
0.04
udef and vnorm predictions due to Ca , Cnom , and the ensemble mean
of “pushed-forward posterior” simulations, and compare them with
experimental observations. As in Figs. 7, 9, and 11, we see a peak in
0.02 udef in the middle of the domain (the jet core) and a large value of the
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
y [m]
Fig. 12 Contour plots of the vorticity field generated on the cross-plane
using Copt (solid line, –) and Ca (dashed line, – –) for the M  0.7; Table 1 Values of Copt obtained for the test cases in this study
J  10.2 test case. We do not have experimental measurements for this
test case. Test case Optimal Cμ ; Cϵ2 ; Cϵ1 
M  0.8; J  10.2 Cga
opt , from Ref. [9] f0.105; 2.099; 1.42g
Note that in this study we do not test the generalizability of a JPDF M  0.8; J  10.2 f0.11; 2.095; 1.44g
M  0.7; J  10.2 f0.11; 2.09; 1.46g
developed for an M; J to other M; J combinations. As mentioned M  0.8; J  16.7 f0.11; 2.04; 1.45g
above, generalization was studied in Ref. [9], with encouraging M  0.6; J  10.2 f0.117; 2.09; 1.2g
results. It led us to believe that a modification of the JPDF for the Ca f0.1; 2.0; 1.34g
M  0.8; J  10.2 could be predictive for other conditions, and
consequently this study. Also, note that in Figs. 7, 9, and 11 we plot Cga is an estimate from Ref. [9], and Ca was derived in this study.
14 Article in Advance / RAY ET AL.

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


16 16 16

14 14 Experiment 14
Ensemble mean
12 12 Cnom 12
Ca
10 10 10
y/dj

y/dj

y/dj
8 8 8

6 6 6

4 4 4

2 2 2

0 0.1 0.2 0.3 0 0.1 0.2 0.3 0 0.1 0.2 0.3


(Umax - u)/U (Umax - u)/U (Umax - u)/U
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


18 18 18
16 16 16
14 14 14
12 12 12
10 10 10
y/dj

y/dj

y/dj
8 8 8
6 6 6 Experiment
Ensemble mean
4 4 4 Cnom
2 2 2 Ca

0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
v/U v/U v/U
Fig. 13 Streamwise velocity deficit udef (above) and normalized vertical velocity vnorm (below) computed using Cnom and Ca compared with the ensemble
mean from the “pushed-forward posterior” simulations and experimental measurements. Results are plotted for the M  0.6; J  16.7 test case, at
three streamwise locations x∕dj  21; 31.5, and 42.0. We see that both Ca and Copt provide better predictions than the nominal parameters Cnom , though
their agreement is not as good as those seen in Figs. 11, 7, and 9.

same in the boundary layer at the bottom of the test section. In VII. Conclusions
between the jet core and the floor of the test section, udef should have Despite their shortcomings, two-equation Reynolds-averaged
decreased to a small number outside the jet; this is seen in Figs. 7, 9, Navier–Stokes (RANS) models, for example, k − ϵ and k − ω
and 11 for all the test cases. As is clear in Fig. 13, the measurements models, are the backbone of engineering aerodynamic simulations.
do not show this trend for the M  0.6; J  10.2 case, though The nominal values of RANS parameters Cnom are not predictive and
predictions using Cnom and Ca do. We cannot explain this aspect of they are often tuned. k − ϵ RANS models have been shown to be not
the experimental measurements, but we do calibrate to it. This could predictive for compressible jet-in-crossflow (JIC) simulations, and in
be the explanation behind the “anomalous” PDF for Cϵ1. Despite the view of their engineering relevance, it is clear that a replacement for
calibration, the k − ϵ RANS model does not reproduce the flowfield Cnom would be useful. It is less clear what that replacement might be,
well (as seen in Fig. 13) leading to large σ in Fig. 4. and whether it would hold across a range of crossflow Mach numbers
Table 1 contains a list of optimal C that we have identified in this and jet strengths. This study has attempted to investigate the problem
study, as well as from previous Bayesian calibrations using rigorously, and may have found a replacement. Note that this rigor
experimental data employed in this study. From Fig. 4 and Table 1, does not endow our estimate of k − ϵ constants with any degree of
we see that Ca is close to the calibrated values, reproduces midplane universality across types of flows—it has been shown to be predictive
measurements well, and is far more predictive than Cnom in all cases for compressible JIC simulations only. Its predictive skill for other
(mid- and cross-plane predictions). Further, Ca respects the trends classes of flows has not been tested.
seen in the JPDFs in Fig. 4—Cμ and Cϵ2 are higher than their nominal In one of our previous papers on this topic [9], a Bayesian method
counterparts, whereas Cϵ1 is similar to the nominal value. In addition, has been developed to compute joint probability density functions
the analytical model from which Ca is derived has been used to (JPDFs) of k − ϵ constants. The JPDF captures the uncertainty in the
predict the trajectory of the jet [13]. These analytical predictions inferred k − ϵ constants. The decision to treat the k − ϵ constants as
have been compared with k − ϵ predictions using Ca , Cnom , and random variables and compute their JPDF from data was deliberate.
experimental data (the M  0.6; 0.7; 0.8, J  10.2 test cases). In The approximation errors in RANS were carefully considered, as
all three cases, both Ca and the analytical trajectory agreed better with well as the sparsity of experimental compressible JIC datasets that
experimental data than k − ϵ RANS predictions using Cnom . It is thus contained measurements of velocity (as opposed to just the trajectory
clear that the analytical model developed in Sec. IV embodies the of the jet), neither of which are conducive to parameter estimation
essential physics in JIC interactions, is far more predictive than Cnom , with a great deal of certainty. As discussed in Sec. I, this work led us
and is close to the calibrated model parameters developed in this to believe that parametric uncertainties, rather than model-form
study. It could serve as an alternative to Cnom for k − ϵ RANS errors, might be responsible for the poor predictive skill of k − ϵ
simulations of compressible JIC interactions. RANS for JIC interactions. Our second paper [11] investigated
Article in Advance / RAY ET AL. 15

whether the inability of RANS models with linear eddy viscosities measured turbulent stresses, as it has been illustrated in one of
to model anisotropy in turbulence could account for the poor our previous papers [9]. The analytical model is entirely derived
predictive power of RANS for JIC interactions. It was found that using mean flow quantities rather than turbulent stresses and thus
nonlinear eddy viscosity models too had to be calibrated to data, and, contributes nothing to rectify this shortcoming. Consequently,
post-calibration, were not vastly more predictive than linear eddy further improvements of the calibrated k − ϵ model for JIC
viscosities. This study reinforced our belief that a modification of simulations will require augmentation of the equations to reduce
Cnom would be sufficient to improve RANS JIC simulations. approximation errors, for example, by employing eddy viscosity
Both the calibration studies mentioned above ran the risk of the models that accommodate anisotropy. Methods and approaches that
inferred JPDFs for k − ϵ constants could compensate for model-form might help in this regard were reviewed in Sec. II. Specifically, the
errors. However, if JPDFs were developed for individual experiments method developed by Ling et al. [36] may serve as an attractive
(and thus compensated for model-form errors to differing degrees), starting point as it has already been applied successfully in
any points of similarity between them could indicate how Cnom would incompressible jet-in-crossflow interactions [38,48].
have to change to be predictive. These JPDFs are plotted in Fig. 4 and
show that 1) Cϵ2 and Cμ would have to be larger than their nominal
counterparts and 2) Cϵ1 could remain unchanged. An analytical Appendix: Checking the Prior for M  0.8; J  16.7
model of an incompressible JIC interaction was also developed, Test Case
and k − ϵ constants’ estimates were obtained. Because this involved In this appendix we investigate the cause of the poor predictions
no model fitting to our datasets, the analytical model is free of of the midplane velocities seen in Fig. 9. We check if the prior Π1 C
can bracket the observations for the M  0.8; J  16.7 test case.
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

any compensation of model-form errors. It was found that these


“analytical” k − ϵ parameters obey the trends seen in the JPDFs. We pick 100 random samples of C from the prior and simulate the flow.
Further flowfields predicted using the JPDFs and the analytical k − ϵ In Fig. A1, we plot the mean prediction from the ensemble of
constants are similar, and in turn are closer to measurements than simulations (solid line) along with the 3 standard deviation bounds
predictions with Cnom . These outcomes lead us to believe that the (dashed lines). The experimental measurements are plotted as
effect of model-form errors on the JPDFs are small, the predictive symbols. We also plot a horizontal line at y∕dj  3. We only use the
skill of Ca would carry over to M; J combinations for which measurements above the horizontal line in our calibration study to
measurements were not available and Ca would be preferable to Cnom remove any influence of the boundary layer on the JPDF of C. We see
for JIC simulations. that the dashed lines do bracket the experimental data, except for about
Having calibrated k − ϵ constants, any residual model–data 10 probes in the last station of probes at x∕dj  42. This shortcoming
mismatch is largely due to model-form errors. This manifests will necessarily impair the quality of the calibration. Note that the
itself most prominently as the disagreement between modeled and 10 probes are a small fraction of the 121 probes used in this study.

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


16 16 16
Experiment
14 14 Ensemble mean 14

12 12 12

10 10 10
y/dj

y/dj

y/dj

8 8 8

6 6 6

4 4 4

2 2 2

0 0.1 0.2 0.3 0 0.1 0.2 0.3 0 0.1 0.2 0.3


(Umax - u)/U (Umax - u)/U (Umax - u)/U

x / dj = 21.0 x / dj = 31.5 x / dj = 42.0


18 18 18
16 16 16
14 14 14
12 12 12
10 10 10
y/dj

y/dj

y/dj

8 8 8
6 6 6
4 4 4
Experiment
2 2 2 Ensemble mean

0 0.2 0.4 0.6 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
v/U v/U v/U
Fig. A1 Streamwise velocity deficit (above) and normalized vertical velocity (below) computed using 100Cnom values sampled from the prior of the
M  0.8; J  16.7 test case, at three streamwise locations x∕dj  21; 31.5, and 42.0). The symbols denote the experimental measurements. The solid
line is the mean prediction of the ensemble of 100 simulations, and the dashed line is the 3ζ bounds of the simulation, where ζ is the standard deviation of
predictions.
16 Article in Advance / RAY ET AL.

Acknowledgments [14] Hasselbrink, E. F., and Mungal, M. G., “Transverse Jets and Jet Flames.
Part 1. Scaling Laws for Strong Transverse Jets,” Journal of Fluid
This work was supported by Sandia National Laboratories’ Mechanics, Vol. 443, Sept. 2001, pp. 1–25.
Advanced Scientific Computing (ASC) Verification and Validation doi:10.1017/S0022112001005146
program. Sandia National Laboratories is a multimission laboratory [15] Broadwell, J. E., and Breidenthal, R. E., “Structure and Mixing of a
managed and operated by National Technology and Engineering Transverse Jet in Incompressible Flow,” Journal of Fluid Mechanics,
Solutions of Sandia LLC, a wholly owned subsidiary of Honeywell Vol. 148, Nov. 1984, pp. 405–412.
International Inc. for the U.S. Department of Energy’s National doi:10.1017/S0022112084002408
Nuclear Security Administration under contract DE-NA0003525. [16] Karagozian, A. R., “An Analytical Model for the Vorticity Associated
with a Transverse Jet,” AIAA Journal, Vol. 24, No. 3, 1986, pp. 429–436.
This paper describes objective technical results and analysis. Any doi:10.2514/3.9285
subjective views or opinions that might be expressed in the paper do [17] Heister, S., and Karagozian, A., “Vortex Modeling of Gaseous Jets in a
not necessarily represent the views of the U.S. Department of Energy Compressible Crossflow,” 24th Joint Propulsion Conference, Joint
or the United States Government. Propulsion Conferences, AIAA Paper 1988-3270, 1988.
doi:10.2514/6.1988-3270
[18] Thies, A. T., and Tam, C. K. W., “Computation of Turbulent
Axisymmetric and Nonaxisymmetric Jet Flows Using the k − ϵ Model,”
References AIAA Journal, Vol. 34, No. 2, 1996, pp. 309–316.
[1] Mahesh, K., “The Interaction of Jets with Crossflows,” Annual Reviews doi:10.2514/3.13065
of Fluid Mechanics, Vol. 45, No. 1, 2013, pp. 379–407. [19] Shirzadi, M., Mirzaei, P. A., and Naghashzadegan, M., “Improvement of
k − ϵ Turbulence Model for CFD Simulation of Atmospheric Boundary
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

doi:10.1146/annurev-fluid-120710-101115
[2] Peterson, C. W., Wolfe, W. P., and Payne, J. L., “Experiments and Layer Around a High-Rise Building Using Stochastic Optimization and
Computations of Roll Torque Induced by Vortex-Fin Interaction,” 42nd Monte Carlo Sampling Technique,” Journal of Wind Engineering and
AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 2004- Industrial Aerodynamics, Vol. 171, No. Suppl. C, 2017, pp. 366–379.
1069, 2004. doi:10.1016/j.jweia.2017.10.005
doi:10.2514/6.2004-1069 [20] Pope, S. B., “An Explanation of the Turbulent Round-Jet/Plane-Jet
[3] Beresh, S. J., Henfling, J. F., Erven, R. J., and Spillers, R. W., Anomaly,” AIAA Journal, Vol. 16, No. 3, 1978, pp. 279–281.
“Penetration of a Transverse Supersonic Jet into a Subsonic doi:10.2514/3.7521
Compressible Crossflow,” AIAA Journal, Vol. 43, No. 2, 2005, [21] Sarkar, S., and Lakshmanan, B., “Application of a Reynolds Stress
pp. 379–389. Turbulence Model to the Compressible Shear Layer,” AIAA Journal,
doi:10.2514/1.9919 Vol. 29, No. 5, 1991, pp. 743–749.
[4] Beresh, S. J., Henfling, J. F., Erven, R. J., and Spillers, R. W., “Turbulent doi:10.2514/3.10649
Characteristics of a Transverse Supersonic Jet in a Subsonic [22] Cheung, S. H., Oliver, T. A., Prudencio, E. E., Prudhomme, S., and
Compressible Crossflow,” AIAA Journal, Vol. 43, No. 11, 2005, Moser, R. D., “Bayesian Uncertainty Analysis with Applications to
pp. 2385–2394. Turbulence Modeling,” Reliability Engineering and System Safety,
doi:10.2514/1.14575 Vol. 96, No. 9, 2011, pp. 1137–1149.
[5] Beresh, S. J., Henfling, J. F., Erven, R. J., and Spillers, R. W., doi:10.1016/j.ress.2010.09.013
“Crossplane Velocimetry of a Transverse Supersonic Jet in a Transonic [23] Edeling, W. N., Cinnella, P., Dwight, R. P., and Bijl, H., “Bayesian
Crossflow,” AIAA Journal, Vol. 44, No. 12, 2006, pp. 3051–3061. Estimates of Parameter Variability in k − ϵ Turbulence Model,” Journal
doi:10.2514/1.22311 of Computational Physics, Vol. 258, Feb. 2014, pp. 73–94.
[6] Beresh, S. J., Heineck, J. T., Walker, S. M., Shairer, E. T., and Yaste, D. doi:10.1016/j.jcp.2013.10.027
M., “Vortex Structure Produced by a Laterally Inclined Supersonic Jet in [24] Edeling, W. N., Cinnella, P., and Dwight, R. P., “Predictive RANS
Transonic Crossflow,” Journal of Propulsion and Power, Vol. 23, No. 2, Simulations via Bayesian Model-Scenario Averaging,” Journal of
2007, pp. 353–363. Computational Physics, Vol. 275, Oct. 2014, pp. 65–91.
doi:10.2514/1.25444 doi:10.1016/j.jcp.2014.06.052
[7] Beresh, S. J., Heineck, J. T., Walker, S. M., Schairer, E. T., and Yaste, D. [25] Guillas, S., Glover, N., and Malki-Epshtein, L., “Bayesian Calibration
M., “Planar Velocimetry of Jet/Fin Interaction on a Full-Scale Flight of the Constants of the k − ϵ Turbulence Model for a CFD Model of
Vehicle Configuration,” AIAA Journal, Vol. 45, No. 8, 2007, pp. 1827– Street Canyon Flow,” Computer Methods in Applied Mechanics and
1840. Engineering, Vol. 279, Sept. 2014, pp. 536–553.
doi:10.2514/1.26485 doi:10.1016/j.cma.2014.06.008
[8] Arunajatesan, S., “Evaluation of Two-Equations RANS Models for [26] Zhang, J., and Fu, S., “An Efficient Bayesian Uncertainty Quantification
Simulation of Jet-in-Crossflow Problems,” 50th AIAA Aerospace Approach with Application to k − ϵ − γ Transition Modeling,”
Sciences Meeting Including the New Horizons Forum and Aerospace Computers & Fluids, Vol. 161, No. Suppl. C, 2018, pp. 211–224.
Exposition, Aerospace Sciences Meetings, AIAA Paper 2012-1199, doi:10.1016/j.compfluid.2017.11.007
2012. [27] Kato, H., Ishiko, K., and Yoshizawa, A., “Optimization of Parameter
doi:10.2514/6.2012-1199 Values in the Turbulence Model Aided by Data Assimilation,” AIAA
[9] Ray, J., Lefantzi, S., Arunajatesan, S., and Dechant, L., “Bayesian Journal, Vol. 54, No. 5, 2016, pp. 1512–1523.
Parameter Estimation of a k-Epsilon Model for Accurate Jet-in- doi:10.2514/1.J054109
Crossflow Simulations,” AIAA Journal, Vol. 54, No. 8, 2016, pp. 2432– [28] Dow, E., and Wang, Q., “Quantification of Structural Uncertainties in
2448. the k − ω Turbulence Model,” 52nd AIAA/ASME/ASCE/AHS/ASC
doi:10.2514/1.J054758 Structures, Structural Dynamics and Materials Conference, Structures,
[10] Wilcox, D. C., Turbulence Modeling for CFD, DCW Industries, La Structural Dynamics, and Materials and Co-Located Conferences,
Canada, CA, 1998, Chap. 4. AIAA Paper 2011-1762, 2011.
[11] Ray, J., Lefantzi, S., Arunajatesan, S., and DeChant, L., “Learning an doi:10.2514/6.2011-1762
Eddy Viscosity Model Using Shrinkage and Bayesian Calibration: A [29] Weatheritt, J., and Sandberg, R., “A Novel Evolutionary Algorithm
Jet-in-Crossflow Case Study,” ASME Journal of Risk Uncertainty Part Applied to Algebraic Modifications of the RANS Stress–Strain
B, Vol. 4, No. 1, 2017, Paper 011001. Relationship,” Journal of Computational Physics, Vol. 325, No. Suppl.
doi:10.1115/1.4037557 C, 2016, pp. 22–37.
[12] Ray, J., Lefantzi, S., Arunajatesan, S., and DeChant, L. J., “Robust doi:10.1016/j.jcp.2016.08.015
Bayesian Calibration of a RANS Model for Jet-in-Crossflow [30] Gorlé, C., and Iaccarino, G., “A Framework for Epistemic Uncertainty
Simulations,” 8th AIAA Theoretical Fluid Mechanics Conference, Quantification of Turbulent Scalar Flux Models for Reynolds-Averaged
AIAA AVIATION Forum, AIAA Paper 2017-4166, 2017. Navier-Stokes Simulations,” Physics of Fluids, Vol. 25, No. 5, 2013,
doi:10.2514/6.2017-4166 Paper 055105.
[13] DeChant, L. J., Ray, J., Lefantzi, S., Ling, J., and Arunajatesan, S., doi:10.1063/1.4807067
“K − ϵ Turbulence Model Parameter Estimates Using an Approximate [31] Wu, J.-L., Wang, J.-X., and Xiao, H., “A Bayesian Calibration–
Self-Similar Jet-in-Crossflow Solution,” 8th AIAA Theoretical Fluid Prediction Method for Reducing Model-Form Uncertainties with
Mechanics Conference, AIAA AVIATION Forum, AIAA Paper 2017- Application in RANS Simulations,” Flow, Turbulence and Combustion,
4167, 2017. Vol. 97, No. 3, 2016, pp. 761–786.
doi:10.2514/6.2017-4167 doi:10.1007/s10494-016-9725-6
Article in Advance / RAY ET AL. 17

[32] Wang, J.-X., Wu, J.-L., and Xiao, H., “Physics-Informed Machine Vol. 45, No. 5, 2007, pp. 1036–1046.
Learning Approach for Reconstructing Reynolds Stress Modeling doi:10.2514/1.21075
Discrepancies Based on DNS Data,” Physical Review Fluids, Vol. 2, [40] Haario, H., Laine, M., Mira, A., and Saksman, E., “DRAM-Efficient
March 2017, Paper 034603. Adaptive MCMC,” Statistics and Computing, Vol. 16, No. 4, 2006,
doi:10.1103/PhysRevFluids.2.034603 pp. 339–354.
[33] Parish, E. J., and Duraisamy, K., “A Paradigm for Data-Driven Predictive doi:10.1007/s11222-006-9438-0
Modeling Using Field Inversion and Machine Learning,” Journal of [41] Raftery, A., and Lewis, S. M., “Implementing MCMC,” Markov Chain
Computational Physics, Vol. 305, No. Suppl. C, 2016, pp. 758–774. Monte Carlo in Practice, edited by W. R. Gilks, S. Richardson, and D. J.
doi:10.1016/j.jcp.2015.11.012 Spiegelhalter, Chapman and Hall, Boca Raton, FL, 1996, pp. 115–130,
[34] Singh, A. P., Medida, S., and Duraisamy, K., “Machine-Learning- Chap. 7.
Augmented Predictive Modeling of Turbulent Separated Flows over [42] R Core Team, R: A Language and Environment for Statistical
Airfoils,” AIAA Journal, Vol. 55, No. 7, 2017, pp. 2215–2227. Computing, R Foundation for Statistical Computing, Vienna, 2012,
doi:10.2514/1.J055595 http://www.R-project.org/.
[35] Edeling, W. N., Iaccarino, G., and Cinnella, P., “Data-Free and [43] Soetaert, K., and Petzoldt, T., “Inverse Modelling, Sensitivity and Monte
Data-Driven RANS Predictions with Quantified Uncertainty,” Flow, Carlo Analysis in R Using Package FME,” Journal of Statistical
Turbulence and Combustion, Vol. 100, No. 3, 2017, pp. 593–616. Software, Vol. 33, No. 3, 2010, pp. 1–28, http://www.jstatsoft.org/
doi:10.1007/s10494-017-9870-6 v33/i03/.
[36] Ling, J., Kurzawski, A., and Templeton, J., “Reynolds Averaged doi:10.18637/jss.v033.i03
Turbulence Modelling Using Deep Neural Networks with Embedded [44] Warnes, G. R., and with Contributions by Robert Burrows, mcgibbsit:
Invariance,” Journal of Fluid Mechanics, Vol. 807, Nov. 2016, Warnes and Raftery’s MCGibbsit MCMC Diagnostic, 2011, http://
Downloaded by UNIVERSITY OF VIRGINIA on November 23, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J057204

pp. 155–166. CRAN.R-project.org/package=mcgibbsit.


doi:10.1017/jfm.2016.615 [45] Silverman, B. W., Density Estimation for Statistics and Data Analysis,
[37] Barone, M. F., Ling, J., Chowdhary, K., Davis, W., and Fike, J., Chapman and Hall, Boca Raton, FL, 1986, Chap. 3.
“Machine Learning Models of Errors in Large Eddy Simulation [46] Pope, S. B., Turbulent Flows, Cambridge Univ. Press, Cambridge, U.K.,
Predictions of Surface Pressure Fluctuations,” 47th AIAA Fluid 2009.
Dynamics Conference, AIAA AVIATION Forum, AIAA Paper 2017- [47] Tennekes, H., and Lumley, J., A First Course in Turbulence, MIT Press,
3979, 2017. Cambridge, MA, 1987, pp. 104–145.
doi:10.2514/6.2017-3979 [48] Ling, J., Ruiz, A., Lacaze, G., and Oefelein, J., “Uncertainty Analysis
[38] Milani, P. M., Ling, J., Saez-Mischlich, G., Bodart, J., and Eaton, J. K., and Data-Driven Model Advances for a Jet-in-Crossflow,” ASME
“A Machine Learning Approach for Determining the Turbulent Journal of Turbomachinery, Vol. 139, No. 2, 2016, Paper 021008.
Diffusivity in Film Cooling Flows,” ASME Journal of Turbomachinery, doi:10.1115/1.4034556
Vol. 140, No. 2, 2017, Paper 021006.
doi:10.1115/1.4038275 R. Ghanem
[39] Brinckman, K. W., Calhoon, W. H., and Dash, S. M., “Scalar Fluctuation Associate Editor
Modeling for High-Speed Aeropropulsive Flows,” AIAA Journal,

You might also like