Evaluation of Turbulence Model Corrections for Supersonic Jets using the OVERFLOW Code

Similar documents
Analysis of Modified Compressibility Corrections for Turbulence Models

Entitled Assessment of Turbulence Models and Corrections with Application to the Orion Launch Abort System

Computational Fluid Dynamics Analysis of Jets with Internal Forced Mixers

Computational Investigations of High-Speed Dual-Stream Jets

NAPC Numerical investigation of axisymmetric underexpanded supersonic jets. Pratikkumar Raje. Bijaylakshmi Saikia. Krishnendu Sinha 1

Numerical Investigation of the Transonic Base Flow of A Generic Rocket Configuration

Numerical Study of Jet Plume Instability from an Overexpanded Nozzle

Large Eddy Simulation of Three-Stream Jets

The Computations of Jet Interaction on a Generic Supersonic Missile

RANS-Based Noise Predictions of Jets with Internal Forced Mixers *

FEDSM COMPUTATIONAL AEROACOUSTIC ANALYSIS OF OVEREXPANDED SUPERSONIC JET IMPINGEMENT ON A FLAT PLATE WITH/WITHOUT HOLE

NUMERICAL INVESTIGATION ON THE FLOW CHARACTERISTICS OF A SUPERSONIC JET IMPINGING ON AN AXI-SYMMETRIC DEFLECTOR

Numerical Studies of Supersonic Jet Impingement on a Flat Plate

Detached Eddy Simulation on Hypersonic Base Flow Structure of Reentry-F Vehicle

compression corner flows with high deflection angle, for example, the method cannot predict the location

AERODYNAMIC SHAPING OF PAYLOAD FAIRING FOR A LAUNCH VEHICLE Irish Angelin S* 1, Senthilkumar S 2

EVALUATION OF TWO-DIMENSIONAL HIGH-SPEED TURBULENT BOUNDARY LAYER AND SHOCK-BOUNDARY LAYER INTERACTION COMPUTATIONS WITH THE OVERFLOW CODE.

Numerical investigation of swirl flow inside a supersonic nozzle

Simulation of unsteady muzzle flow of a small-caliber gun

CFD Performance of Turbulence Models for Flow from Supersonic Nozzle Exhausts

LEE-SIDE FLOW SIMULATIONS OF CRUCIFORM WING- BODY CONFIGURATIONS AT INCOMPRESSIBLE MACH NUMBERS

Laminar and Turbulent Flow Calculations for the HIFiRE-5b Flight Test

A comparison of turbulence models for an impinging jet in a crossflow

Numerical Methods in Aerodynamics. Turbulence Modeling. Lecture 5: Turbulence modeling

Numerical Study of Pressure Waves Generated by H-IIA Launch Vehicle at Lift-off

WALL ROUGHNESS EFFECTS ON SHOCK BOUNDARY LAYER INTERACTION FLOWS

Numerical Simulations of the Mars Science! Laboratory Supersonic Parachute!

Computational Analysis of Bell Nozzles

DIRECT NUMERICAL SIMULATIONS OF HIGH SPEED FLOW OVER CAVITY. Abstract

The Simulation of Wraparound Fins Aerodynamic Characteristics

Numerical Simulation of Jet Aerodynamics Using the Three-Dimensional Navier-Stokes Code PAB3D

41st Aerospace Sciences Meeting and Exhibit 6-9 January 2003, Reno, Nevada Modeling shock unsteadiness in shock/turbulence interaction

Experimental investigation of flow control devices for the reduction of transonic buffeting on rocket afterbodies

There are no simple turbulent flows

Turbulence Model Effects on Cold-Gas Lateral Jet Interaction in a Supersonic Crossflow

REDUCTION OF AERODYNAMIC HEATING AND DRAG WITH OPPOSING JET THROUGH EXTENDED NOZZLE IN HIGH ENTHALPY FLOW

Numerical investigation on the effect of inlet conditions on the oblique shock system in a high-speed wind tunnel

Prospects for High-Speed Flow Simulations

The Very Near Pressure Field of Three-Stream Jets

DESIGN & COMPUTATIONAL FLUID DYNAMICS ANALYSES OF AN AXISYMMETRIC NOZZLE AT TRANSONIC FREE STREAM CONDITIONS

Eigenmode Analysis of Boundary Conditions for the One-dimensional Preconditioned Euler Equations

The effect of geometric parameters on the head loss factor in headers

EFFECT OF REYNOLDS NUMBER ON THE UNSTEADY FLOW AND ACOUSTIC FIELDS OF SUPERSONIC CAVITY

Resolving the dependence on free-stream values for the k-omega turbulence model

A Computational Investigation of a Turbulent Flow Over a Backward Facing Step with OpenFOAM

CHARACTERISTICS OF ELLIPTIC CO-AXIAL JETS

URANS Computations of Cavitating Flow around a 2-D Wedge by Compressible Two-Phase Flow Solver

AIAA Computational Analysis of a Pylon-Chevron Core Nozzle Interaction

Assessment of Implicit Implementation of the AUSM + Method and the SST Model for Viscous High Speed Flow

Studies on the Transition of the Flow Oscillations over an Axisymmetric Open Cavity Model

CFD Analysis of Micro-Ramps for Hypersonic Flows Mogrekar Ashish 1, a, Sivakumar, R. 2, b

TURBULENT FLOW ACROSS A ROTATING CYLINDER WITH SURFACE ROUGHNESS

Comparison of Turbulence Models for Nozzle-Afterbody Flows With Propulsive Jets

INVESTIGATION OF FLOW PARAMETERS AND NOISE OF SUBSONIC AND SUPERSONIC JETS USING RANS/ILES HIGH RESOLUTION METHOD

Manhar Dhanak Florida Atlantic University Graduate Student: Zaqie Reza

STUDY OF THREE-DIMENSIONAL SYNTHETIC JET FLOWFIELDS USING DIRECT NUMERICAL SIMULATION.

All-Particle Multiscale Computation of Hypersonic Rarefied Flow

Elliptic Trailing Edge for a High Subsonic Turbine Cascade

Verification of K-ω SST Turbulence Model for Supersonic Internal Flows

Publication 97/2. An Introduction to Turbulence Models. Lars Davidson, lada

Hypersonic Turbulent Flow Simulation of FIRE II Reentry Vehicle Afterbody

Aeroacoustics, Launcher Acoustics, Large-Eddy Simulation.

On the transient modelling of impinging jets heat transfer. A practical approach

NUMERICAL SIMULATION OF HIGH-SPEED SEPARATION FLOW IN THE AEROSPACE PROPULSION SYSTEMS

SIMULATION OF THREE-DIMENSIONAL INCOMPRESSIBLE CAVITY FLOWS

Lecture1: Characteristics of Hypersonic Atmosphere

Application of Compact Schemes to Large Eddy Simulation of Turbulent Jets

ASYMMETRY EFFECTS IN NUMERICAL SIMULATION OF SUPERSONIC FLOWS WITH UPSTREAM SEPARATED REGIONS

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Planning for a Supersonic Retropropulsion Test in the NASA Langley Unitary Plan Wind Tunnel

Turbulent Boundary Layers & Turbulence Models. Lecture 09

Numerical Simulation of Supersonic Expansion in Conical and Contour Nozzle

AA214B: NUMERICAL METHODS FOR COMPRESSIBLE FLOWS

In the continuing effort to better understand hypersonic flow phenomena, mixing characteristics of various

Calculations on a heated cylinder case

White Paper FINAL REPORT AN EVALUATION OF THE HYDRODYNAMICS MECHANISMS WHICH DRIVE THE PERFORMANCE OF THE WESTFALL STATIC MIXER.

Analysis of Shock Motion in STBLI Induced by a Compression Ramp Configuration Using DNS Data

A Non-Intrusive Polynomial Chaos Method For Uncertainty Propagation in CFD Simulations

Transition Modeling Activities at AS-C²A²S²E (DLR)

Heating effects on the structure of noise sources of high-speed jets

Investigation of Flow Field in a Typical Hypersonic Wind Tunnel over a Standard Mode

Development of Flow over Blunt-Nosed Slender Bodies at Transonic Mach Numbers

CFD ANALYSIS OF CD NOZZLE AND EFFECT OF NOZZLE PRESSURE RATIO ON PRESSURE AND VELOCITY FOR SUDDENLY EXPANDED FLOWS. Kuala Lumpur, Malaysia

Shape Optimisation of Axisymmetric Scramjets

SUPERSONIC JET CONTROL WITH INTERNAL GROOVES

Semi-Empirical Prediction of Noise from Non-Zero Pressure Gradient Turbulent Boundary Layers

Zonal hybrid RANS-LES modeling using a Low-Reynolds-Number k ω approach

TURBULENCE MODELING VALIDATION FOR AFTERBODY FLOWS A. Hadjadj 1 and A.N. Kudryavtsev 2

A finite-volume algorithm for all speed flows

VERTICAL TURBULENT BUOYANT HELIUM JET CFD MODELING AND VALIDATION

Detailed Outline, M E 320 Fluid Flow, Spring Semester 2015

Simulation of Aeroelastic System with Aerodynamic Nonlinearity

ON USING ARTIFICIAL COMPRESSIBILITY METHOD FOR SOLVING TURBULENT FLOWS

Review and Assessment of Turbulence Transition Models

AN UNCERTAINTY ESTIMATION EXAMPLE FOR BACKWARD FACING STEP CFD SIMULATION. Abstract

Wall treatments and wall functions

Comparison of two equations closure turbulence models for the prediction of heat and mass transfer in a mechanically ventilated enclosure

arxiv: v1 [physics.flu-dyn] 11 Oct 2012

Over-expansion Effects on Mach 3.0 Supersonic Jet Acoustics

COMPUTATIONAL STUDY OF SEPARATION CONTROL MECHANISM WITH THE IMAGINARY BODY FORCE ADDED TO THE FLOWS OVER AN AIRFOIL

Transcription:

Evaluation of Turbulence Model Corrections for Supersonic Jets using the OVERFLOW Code Nico Gross, Gregory A. Blaisdell, and Anastasios S. Lyrintzis Purdue University, West Lafayette, IN 47907 The OVERFLOW code is used for computations of supersonic jets. Corrections for compressibility and temperature effects are evaluated based on comparisons with experimental data. Test cases include isothermal and heated, perfectly expanded and underexpanded axisymmetric jets with design Mach numbers around two. Also, a 3-D jet in an angled crossflow at a design Mach number of three is evaluated. The SST turbulence model is chosen for the flows examined in this study. The Sarkar, Zeman, and Wilcox compressibility corrections and Abdol-Hamid s temperature correction are assessed when used with the SST model. Results for the axisymmetric jets show that using a compressibility correction is important for a good prediction of the jet development. The effect of the temperature correction is not as significant. For the 3-D jet in crossflow, conclusions are not as decisive. The velocity decay is underpredicted when using the compressibility correction. The temperature correction behaves as it did for the axisymmetric jet. The overall recommendation is that the corrections must be used with caution, especially when approaching high temperature and Mach number regimes. Nomenclature a = speed of sound CC = compressibility correction C µ = eddy viscosity coefficient D = nozzle exit diameter ε d = dilatation dissipation ε s = solenoidal dissipation k = kinetic energy of turbulence fluctuations M = Mach number M t = turbulence Mach number PR = pressure ratio R = radial (crossflow) coordinate R 1/2 = jet half radius T = jet temperature TC = temperature correction TR = temperature ratio T inf = ambient temperature T g = temperature gradient U = jet velocity in X-direction U jet = jet centerline velocity in X-direction at nozzle exit plane Vel = jet velocity magnitude X = axial (streamwise) coordinate γ = specific-heat ratio ω = specific dissipation rate Graduate Research Assistant, School of Aeronautics and Astronautics, Student Member AIAA. Associate Professor, School of Aeronautics and Astronautics, Associate Fellow AIAA. Professor, School of Aeronautics and Astronautics, Associate Fellow AIAA. 1 of 21

I. Introduction OVERFLOW 1 is a computational fluid dynamics (CFD) code used for the prediction of aerodynamics and heat transfer of the Space Shuttle and future Constellation launch vehicles. The work presented here is carried out to support computational analysis of the Orion spacecraft, which will bring humans into orbit, to the moon, and eventually to Mars. A key component of the spacecraft is the Launch Abort System (LAS), which is intended to separate the Crew Module (CM) from the rest of the rocket in case of an emergency during ascent. The LAS consists of the four Abort Motors (AM), four Jettison Motors (JM), and eight Attitude Control Motors (ACM), as shown in Fig. 1. Understanding the interaction between the rocket exhaust plumes, the freestream, and the capsule is important in the design and analysis for the spacecraft. For example, an impinging jet might create large localized pressure and temperature peaks. In addition, the shock cells caused by an underexpanded jet present near the capsule surface may cause large changes in pressure coefficients. All these characteristics will impact the flight mechanics of the vehicle. Figure 1. Diagram of Launch Abort System (LAS) showing Abort Motors (AM), Jettison Motors (JM), and Attitude Control Motors (ACM). Courtesy of NASA 2. In fluid mechanics, jets represent the viscous and turbulent interaction between a coherent stream exiting a nozzle and the surrounding freestream. The shear layers, which form at the nozzle boundary layer, mix with the surrounding fluid. The center of the jet, known as the potential core, remains essentially inviscid until the shear layers meet. After this point, the jet becomes fully developed and shows a constant decay in velocity. The velocity profile in the fully developed region is known to be self similar. Parameters of interest when studying jets include the extent of the potential core, the spread rate, and the velocity decay rate. When the jet is no longer axisymmetric, three-dimensional effects must be considered and the jet structure becomes more complex. For a jet in crossflow, the interaction with the freestream causes various vortex formations. For example, the freestream moving around the jet causes the formation of counter-rotating vortex pairs in the main jet stream, and wake vortex interactions downstream. A lot of previous work investigating jet flows has been undertaken. Eggers 3 performed some of the first baseline experiments to examine mean-flow and turbulence quantities in supersonic jets. Birch and Eggers 4 compiled many of the early experiments performed for free shear flows into the Langley curve, which is commonly used to calibrate turbulence models. It was found that the spread rate in mixing layers decreases with increasing Mach number when compressibility effects are present. Lau 5 and Seiner et al. 6 characterized the effects of exit Mach number and temperature on the behavior of jets. They found that an increased jet total temperature enhanced the mixing in the jet due to more density fluctuations. When modeling jets using Reynolds-averaged Navier Stokes (RANS) solvers, turbulence models are critically important. Choosing a turbulence model determines the important parameters in the jet, such as the location of the shocks and the extent of the potential core. Many factors, including the jet Mach number, total temperature ratio, and species composition, can affect CFD predictions. Sarkar et al. 7 showed that additional terms are needed in the turbulence model to capture the decrease in growth rate with increasing Mach number. Georgiadis et al. 8 examined turbulence models specifically developed for jet flows and found that all models showed a deficiency in the initial jet mixing rate and the turbulence kinetic energy predictions. Abdol-Hamid et al. 9 devised a new turbulence model aimed at computations for heated jets. They determined that an additional convective heat transfer model is needed to accurately predict jets with a strong temperature gradient. The objective of the current study is to evaluate the performance of the turbulence model corrections for compressible and heated supersonic flows using OVERFLOW. The jets analyzed are isothermal and heated, and perfectly expanded and underexpanded. The primary variables of interest are the velocity and temperature profiles downstream of the nozzle exit, from which the jet can be characterized. Assessing the code s capability to capture the expansion and compression waves present in an underexpanded jet is also of interest. 2 of 21

II. Test Cases In order to assess OVERFLOW s ability to model supersonic jets, several test cases are computed. Various axisymmetric jet experiments are selected. Then, a 3-D jet in angled crossflow is used to assess further parameters of interest which could not be studied using axisymmetric jets. Experimental data measured using particle image velocimetry (PIV) and pressure probes in the flow field is available for comparison to CFD results. The experiments are chosen based on the quality of the data and the parameters of interest. Key conditions for the test cases considered are summarized in Table 1. Table 1. Reference conditions for test cases considered Test Case Classification M nozzle P 0 /P 0, T 0 /T 0, Eggers isothermal, perfectly expanded 2.2 11.0 1.0 Wishart heated, perfectly expanded 2.0 7.6 2.3 Wishart heated, underexpanded 2.0 9.5 2.3 Glenn heated, underexpanded 3.0 28.5 2.3 A. Eggers Jet The Eggers 3 jet refers to a 1966 experiment characterizing supersonic turbulent jets. A Mach 2.22 isothermal jet with nozzle exit diameter of 1.01 (0.026 m) was exhausted into ambient air, with a pressure ratio of 11. The experiment supplies velocity profiles in the near- and far-field of the jet. This experiment provides a standard test case for supersonic jet validation. B. Wishart Jet The Wishart 10 experiment, performed at Florida State University, is a similar study to that of Eggers, but with conditions more closely resembling those of the LAS. The jet used a nozzle designed for Mach 2, a nozzle exit diameter of 1.15 (0.029 m), and was operated at ideal and off-design conditions. All cases were run isothermal and heated. These conditions allow for examination of the shock-cell structure and temperature effects on the development of the jet. The cases of interest for this study are the heated jet (temperature ratio 2.3) for perfectly expanded (pressure ratio 7.65) and underexpanded (pressure ratio 9.3) conditions. The complete set of computational analysis performed for this experiment can be found in the author s Master s thesis 11. C. Glenn Jet The Glenn jet experiment (Wernet et al. 12 ), performed at the NASA Glenn Research Center, was used to acquire data for a hot supersonic jet in subsonic crossflow. It was performed as part of the aerodynamics analysis and design process for the LAS. While actual LAS flight conditions were not attainable, the jet conditions used here were at the highest possible pressures and temperatures attainable in the test facility, which were stagnation pressure and temperature of 410 psi (2.8 10 6 Pa) and 1350 R (750 K), respectively. A survey of available literature showed that no other experiments had been performed at similar conditions. The nozzle was placed in a wind tunnel with cold freestream air at Mach 0.3. The jet air was heated using a natural gas combustor and exhausted through the conical nozzle at an angle of 25 degrees relative to the freestream. The nozzle exit design Mach number was 3.0, but due to the underexpansion, Mach numbers exceeding four are seen downstream. Experimental data were obtained with PIV at various locations downstream of the nozzle exit and normal to the freestream, producing two dimensional velocity vector fields. III. Numerical Methods A. Code Description The OVERFLOW code 1 is an overset grid solver developed by NASA. It solves the time-dependent, Reynoldsaveraged Navier-Stokes equations for compressible flows. The overset capability makes it useful for computing flow 3 of 21

fields involving complex geometries. The code is formulated using a finite difference method and has various central and upwind schemes for spatial discretization built in. A diagonalized, implicit approximate factorization (ADI) scheme is used for time advancement. Other capabilities, such as local time-stepping and grid sequencing, are also available to accelerate convergence. B. Turbulence Models and Corrections Choosing the correct turbulence model is critical for a good computational prediction of the flowfield. The OVER- FLOW code contains various algebraic, one-equation, and two-equation turbulence models. This study will examine four of the most popular models: the Spalart-Allmaras (SA) model 13, Baldwin-Barth (BB) model 14, k-ω model 15 (1988 version), and Shear Stress Transport (SST) model 16. All turbulence models use the Boussinesq approximation to incorporate the contribution of the Reynolds stresses through an eddy viscosity into the Navier-Stokes equations. The SA and BB models are one-equation models which solve for the eddy viscosity, while the k-ω and SST solve two partial differential equations for the turbulence kinetic energy, k, and the specific dissipation rate, ω. The SST model is a hybrid between the k-ɛ and k-ω models, where a function is used to blend between the models based on distance to the nearest wall. Current turbulence models have been designed for low-speed, isothermal flows. Future research should strive to develop a new model that is general for all types of flows. However, at this time, the practical approach is to modify existing models for more complicated flows. The intent of this study is to assess the various corrections for turbulence models that are applicable to the LAS jet flows, namely compressibility and high temperature corrections. The compressibility correction is devised to deal with additional effects seen for higher Mach number flows, specifically, the effects of compressibility on the dissipation rate of the turbulence kinetic energy. For free shear flows, this is exhibited as the decrease in growth rate in the mixing layer with increasing Mach number 4. Standard turbulence models do not account for this Mach number dependence, and thus a compressibility correction is used. For compressible flows, two extra terms, known as the dilatation dissipation, ε d, and the pressure-dilatation occur in the turbulence kinetic energy equation. The pressure-dilatation term is usually neglected because its contributions have been shown to be small 15. The dilatation dissipation term is included in addition to the solenoidal, or incompressible, dissipation, ε s. In the formulation of the k-ω and SST models, the extra term also occurs in the specific dissipation rate equation. Thus, the effect is that the growth rate of turbulence is inhibited when the correction is active. Sarkar 7 modeled the ratio of the dilatation dissipation to the solenoidal dissipation, ε d /ε s, as a function of the turbulence Mach number, M t, defined as For Sarkar s model, M t = 2k a (1) ε d ε s = F (M t ) = M 2 t (2) The model proposed by Zeman 17 is formulated as follows { F (M t ) = 1 exp [ 12 ]} (γ + 1)(M t M t0 ) 2 /Λ 2 H(M t M t0 ) (3) where M t0 = 0.10 2/(γ + 1) and H(x) is the Heaviside function. Finally, Wilcox s 15 model is given by F (M t ) = (M 2 t M 2 t0)h(m t M t0 ) (4) The model constant M t0 = 0.25. For a complete discussion of the compressibility corrections, see Wilcox 15. It must also me noted that the implementation of the SST turbulence model in OVERFLOW follows that by Suzen and Hoffman 18, where the compressibility correction is multiplied by the SST blending function, thus effectively disabling the correction in near wall regions. Abdol-Hamid et al. 9 noticed that standard turbulence models also fail to capture the increase in the shear layer growth rate due to temperature effects observed by Lau and Seiner et al. They devised a new correction to deal with these effects. The model was calibrated to the supersonic jet experiment of Seiner et al. 6. The correction calculates a modified eddy viscosity coefficient, C µ, based on the normalized total temperature gradient, T g, in the flow, 4 of 21

C µ = 0.09 [ 1 + T 3 g 0.041 + f(m t ) ] (5) F (M t ) is the same as in Eq. 4 above, except the model constant M t0 = 0.1. From this formulation, a large total temperature gradient causes an increase in the eddy viscosity, which thus increases the diffusion in the jet. The correction also considers the compressibility effects by including the F (M t ) term. For a large turbulence Mach number, the eddy viscosity coefficient would be decreased and would suppress the mixing. The compressibility effects in the temperature correction s F (M t ) term are calculated separately from those of an implicit compressibility correction such as Sarkar s. However, for Abdol-Hamid s supersonic jet test case 9, all analysis is completed while also using Sarkar s compressibility correction, and this approach is followed for this work as well. C. Grid Generation Grids for the Eggers and Wishart jets are very similar. The grid for the axisymmetric jets is shown in Fig. 2. Four zones are used for the nozzle, top, aft, and the nozzle lip regions. The X-axis represents the jet s axis of symmetry. The collar grid around the nozzle lip is used to better resolve the high flow gradients present in this region. The grid spacing along the inner nozzle boundary is designed to keep the wall y + below one in order to adequately resolve the viscous sublayer near the wall. The grid is particularly fine in the aft region off the nozzle lip to resolve the shear layer that determines the jet development. In order to sufficiently model the freestream and eliminate upper and aft boundary effects, the domain of the plume extends at least 30 diameters in the radial and 80 diameters in the axial directions, measured from the nozzle exit plane. The final Eggers and Wishart grid systems contained 91.3 10 3 and 105.8 10 3 grid points, respectively. For the Glenn jet, grid scripts were obtained from researchers at NASA Ames Research Center. The model includes the entire wind tunnel. Similar to the grids for the axisymmetric jets, particular attention is paid to ensure sufficient resolution along the nozzle lip and the mixing region. For the final grid, the plume region contained nine million grid points, leading to a total of 29 10 6 grid points. Figure 3 shows the grid for the Glenn jet along the centerline plane. D. Boundary Conditions For the axisymmetric jet, freestream boundary conditions are imposed at the top of the nozzle, with Mach 0.01 and standard pressure and temperature. For the Glenn jet, the freestream boundary conditions are imposed at the tunnel entrance with Mach 0.3. All walls are specified as viscous and adiabatic. The nozzle conditions are set as a ratio of stagnation pressure and temperature from the nozzle plenum to the freestream. For the Glenn jet, the natural gas combustion products in the nozzle are modeled as a second species with specific heats ratio, γ, of 1.32. The top and aft domain boundaries are specified as pressure outflow boundaries with a given value of static pressure. E. Solution Method The numerical schemes used are chosen based on robustness and rate of convergence. For consistency, all test cases use the same inputs for the numerical methods. The HLLC upwind scheme proved better than the central difference and Roe s upwind scheme for the discretization of the spatial derivatives. This is especially true for cases in which shocks are present, as the HLLC is an improved algorithm for capturing shocks with low smearing. A van Albada limiter is used for the upwind Euler terms, along with a MUSCLE scheme flux limiter for added smoothing. The spatial differencing for all convective terms uses third order accuracy. Local timestep scaling with a constant CFL number is employed, where the timestep scaling is based on the local cell Reynolds number. All cases are run with three levels of grid sequencing for improved convergence. 1,000 iterations are run on the coarse and medium grid levels each before running 6,000 iterations on the fine grid. The solution is considered converged when all residuals drop at least three orders of magnitude and momentum forces at various downstream locations are constant. All calculations for the axisymmetric jet are run on a 128 processor cluster at NASA JSC consisting of 64 1.3 GHz and 64 1.5 GHz processors and 124 GB RAM. Runs are split between 20 processors and use an average of 10 CPU hours for the final grid. For the 3-D jet, cases are run on the NASA Advanced Supercomputing (NAS) cluster Columbia using 64 CPUs, with an average CPU time between 1500 and 1800 hours, depending on the case. A small (13%) increase in computation time is observed when using the temperature correction. 5 of 21

F. Grid Convergence Free shear flows such as this jet flow are sensitive to the computational grid. A grid that is too coarse may show an overly diffusive behavior. Grid refinement is thus performed to ensure that the solution is grid independent. Results are shown only for the underexpanded jet as this case showed the most sensitivity to the grid. The grid is refined in the regions of interest: the end of the nozzle and plume where mixing with the freestream occurs. The number of grid points in each direction is doubled successively. The centerline velocity profiles downstream are sensitive to grid refinement and important to the analysis. Figure 4 shows the centerline Mach number for three levels of grid refinement, with 12.1 10 3, 48.5 10 3, and 196.0 10 3 grid points in the plume region. The most significant distinction is seen in the resolution of the shock cells. While all grids predict the length of the shock cells equally, the peak is smeared with the coarsest grid, where the peak Mach number is only 2.58, compared with 2.72 and 2.87 for the medium and fine grids, respectively. After the potential core, there is very little difference in the results. It is decided that the medium resolved all parameters of interest sufficiently, and thus is used in all subsequent computations. A similar study (not presented here) showed that the grid for the 3-D jet is also grid independent. IV. Results A. Eggers Jet The goal of the Eggers jet runs is to evaluate the baseline turbulence model performance. Figure 5a shows the jet centerline velocity, normalized by the jet exit velocity, as a function of downstream distance, normalized by jet exit diameter, for various uncorrected turbulence models. The velocity remains constant through the jet potential core, and then decays in the fully developed region. The SST, k-ω, and SA models predict too much turbulent mixing, thus showing a potential core length that is shorter than that of the experiment. The BB model, on the other hand, significantly underpredicts the mixing and exhibits an excessively long potential core. The SST model gives the best prediction, especially further downstream. Figure 5b shows the jet half radius for the turbulence models. The half radius is defined as the radius at which the velocity is 50% of the local centerline velocity, computed at all downstream (X) locations. This gives a representation of the jet s spreading rate. It can be seen that because of the increased mixing by the SST, k-ω, and SA models, the spread rate is also overpredicted. The BB model does not show enough mixing and does not spread as quickly. The SST model matches the experimental spread rate most closely. Based on this preliminary analysis, the SST turbulence model is chosen for all subsequent computations. Figure 6 compares the compressibility corrections by Sarkar, Zeman, and Wilcox, which were implemented into the SST model in the OVERFLOW source code. All three show a noteworthy improvement over not using any correction. Wilcox s model diffuses slightly more quickly compared with the experimental data. There is no significant difference between the Sarkar and Zeman models. In Fig. 6b, the spread rate for the three compressibility corrections is presented. Analogous to the centerline velocity, the Sarkar and Zeman corrections show the best performance, and all corrections show a significant improvement over the uncorrected model. There is no major difference between the three compressibility corrections examined, and hence Sarkar s model will be used for all further analysis. B. Heated Wishart Jet Figure 7a presents the centerline Mach number for the heated Wishart jet. The plot shown contains the baseline turbulence model (CC=0, TC=0), the model with Sarkar s compressibility correction (CC=1, TC=0), and the model with Abdol-Hamid s temperature correction and Sarkar s compressibility correction (CC=1, TC=1). As observed with the Eggers jet, the compressibility correction decreases turbulent mixing and shifts the curve to the right. The temperature correction increases mixing when used with the compressibility correction. It is important to note that the compressibility correction has the more pronounced effect on the solution. While the use of no corrections most closely matches the experimental potential core length, the downstream behavior is best predicted when using the compressibility correction. In Fig. 7b, the temperature profile, normalized by the ambient temperature, is plotted along the centerline. The increase in temperature is caused by the compression as the gas decelerates in the shear layer, and then decreases to the ambient value through mixing with the freestream. All models show an identical magnitude in the temperature rise, which matches well with the experimental data. Similarly to before, the start of the shear layer is better predicted 6 of 21

without correction, whereas the downstream behavior agrees more closely with the compressibility correction. In both plots, it is indiscernible whether adding the temperature correction provides a better prediction. Figure 8 shows radial Mach number profiles for four downstream locations. The difference in jet development prediction is clearly demonstrated here. There is no noteworthy distinction between the models for X/D of 3.5 and 7 as the jet has not had ample time to mix with the ambient fluid. The compressibility correction again is seen to inhibit mixing, thus showing a longer potential core. For R/D larger than 0.8, all models show a similar prediction as the jet has dissipated into the ambient fluid. For the two later X/D locations, the experimental data lie between the CFD results, and there is no clear model that exhibits a better behavior. However, at increasing radial distance, the model with the compressibility corrections seems to show better agreement with the experimental data. C. Heated and Underexpanded Wishart Jet The contour plots of Mach number in Fig. 9 illustrate the behavior of the underexpanded and heated jet. The jet exit pressure is higher than the back pressure and expansion fans are formed at the lip of the jet. Once the expansion waves reach the jet boundary, they reflect as compression waves to match the constant pressure condition at the boundary. When the pressure ratio is large enough, the compression waves no longer meet at the axis but instead form a stronger normal shock known as the Mach disc. The convergence of the expansion fan generates a cylindrical shock, known as the barrel shock, which terminates at the Mach disc. This leads to a pattern of shock cells, frequently seen in the exhaust of a jet or rocket engine. The wavelength of the expansion and compressions depends on the Mach number and pressure ratio. For this case, the maximum Mach numbers are approaching three, much higher than the nozzle design Mach number of two. It is evident that the choice of turbulence model correction significantly impacts the development of the shock cells. Using exclusively the compressibility correction (b) shows seven distinguishable shock cells, compared with only four for the uncorrected model (a). The use of the temperature correction does not change the number of shock cells present, but does slightly decrease the length when used with the compressibility correction (c). Figures 10a and b show the centerline Mach number and temperature profiles, respectively, for the heated and underexpanded jet. It is noted that the sharp increases in Mach number (due to expansion fans) correspond to sharp decreases in temperature, as expected, and the opposite is true for the compressions. The difference in the number of shock cells present is also visible between the corrections. For X/D less than seven, all models match the experimental data very well. The models without the compressibility correction then decay too rapidly. The SST model with the compressibility correction (and the temperature correction enabled or disabled), matches the data very well, except for the slightly excessive potential core length. As before, the temperature and Mach number profiles correspond, and the compressibility-corrected models show excellent agreement with experimental data. Again, the magnitude of the temperature peak is predicted equally well by all models. When using the temperature correction with the compressibility correction, the temperature correction increases mixing. The radial Mach number profiles for the heated and underexpanded jet are shown in Fig. 11. The Mach number in the potential core is no longer uniform because the expansions and contractions change the flow direction. The compressibility-corrected models show an excellent prediction for all downstream locations. The uncorrected models clearly diffuse too quickly, both in the radial and axial directions. A possible explanation for why the compressibility correction performs better for this case than for the perfectly expanded case is that compressibility effects become more significant in this case because higher Mach numbers are seen due to expansions and contractions. D. Parametric Study A parametric study is performed using a generic axisymmetric jet setup. The goal is to extrapolate the data to higher pressure ratios (PR) and temperature ratios (TR), more closely resembling those experienced during actual LAS flight conditions. The pressure ratio is increased from the perfectly expanded conditions (PR = 11) to underexpanded ratios of 15, 25, and 35. The temperature ratio is successively increased from isothermal to ratios of 5 and 10. The compressibility and temperature corrections are examined. A survey of literature showed that there is very little experimental data available for verification at these conditions. Fig. 12a shows the jet half radius for increasing pressure ratios with the compressibility correction disabled (solid line) and enabled (dashed line). Initially, the half radius is largest for the highest pressure ratio. This is occurs because the higher pressure ratio causes the flow to expand more closer to the nozzle lip, forming the first barrel shock. There is basically no distinction between the different pressure ratios without the compressibility correction after X/D of 7 of 21

20. This is inconsistent with the observation that growth rate should decrease at higher Mach numbers and highlights the deficiency of the base turbulence models. For the compressibility correction, as expected, the growth rate for the higher pressure ratio is decreased. This tends to imply that that the compressibility correction is required to match the flow physics. Fig. 12b illustrates the jet half radius for increasing temperature ratios when the temperature correction is disabled (solid line) and enabled (dashed line). In both cases, the compressibility correction remains enabled, as this is how the temperature correction is intended to be used. The increasing temperature ratios lead to an increase in turbulence kinetic energy. From Eq. 1, this increase causes a rise in the turbulence Mach number. In turn, the turbulence dissipation is increased and the turbulent mixing is decreased, showing the the smaller spread rate seen in the figure. The use of the temperature correction essentially returns the spread rate to the initial rate, offsetting the effect of the compressibility correction. Unfortunately, there is no experimental data at these elevated test conditions, so it cannot be determined which correction accurately models the flow physics. This study does, however, warrant that caution must be exercised when using the corrections at conditions for which they have not been validated. E. Glenn Jet Finally, comparisons are made for the 3-D Glenn jet with crossflow. Figure 13 shows the velocity magnitude contours at four downstream locations perpendicular to the freestream, normalized by the jet exit diameter, and compared with PIV data. Fig. 14 shows cuts through the middle of the velocity contours to give velocity profiles. For the uncorrected model, the velocity magnitude and vortex roll-up diffuse too quickly when compared with experiment. As seen for the axisymmetric analysis, the compressibility correction shows a large effect on the problem. The velocity magnitude is too high at downstream locations because turbulent mixing in the shear layer is suppressed for too long. The compressibility correction also shows increased vorticity and thus a shape that is skewed excessively by the vortex roll-up. From this test case, it appears that Sarkar s model overcorrects by not showing enough mixing with the freestream. A potential reason for this behavior, when compared with the favorable behavior for the axisymmetric jet, is that the current experiment is performed at much higher turbulent Mach numbers than the previous cases. The temperature correction slightly increases the mixing (as before), but does not have a significant effect on the solution. Again, the compressibility correction is dominant for this problem. This is expected because compressibility effects are large for this flow due to high turbulence Mach numbers in the jet shear layer. A. Summary of Key Results V. Conclusions and Future Work The following summarizes key conclusions and recommendations drawn from the previous analysis: The SST model is the best choice out of the turbulence models tested for this flow. The k-ω and SA models are too diffusive, while the BB model suppresses mixing excessively. The use of a compressibility correction is essential for a good prediction of the axisymmetric supersonic jets. While slightly inhibiting mixing in the initial region and thus overpredicting the potential core length, the behavior in the fully developed region is drastically improved. This improvement is most evident in the accurate prediction of the shock cell structure for the cases with an underexpanded jet. The behavior of the temperature correction is not as significant compared to the compressibility correction. For flows in which compressibility effects are significant, the temperature correction should only be used in conjunction with the compressibility correction (as intended when it was formulated by Abdol-Hamid). Both the compressibility and temperature corrections must be further evaluated before being used for higher temperature ratios or Mach numbers. The parametric study shows that the effect of the compressibility and temperature corrections becomes more pronounced at elevated pressure and temperature ratios. The corrections have not been validated against experimental data at these conditions. The 3-D jet with crossflow presents a much more complex flowfield. The physics that must be modeled include various vortex interactions between the jet and crossflow. An example is the formation of counter-rotating vortices which leads to further additional instabilities in the flowfield. 8 of 21

The compressibility correction by Sarkar does not exhibit sufficient mixing at the higher Mach numbers seen for the 3-D jet because it suppresses turbulence growth excessively. For the 3-D jet, the temperature correction behaves in a manner consistent with the previous axisymmetric jet test cases. Again, it does not have as pronounced an effect as the compressibility correction. B. Future Work Examining free shear flows at higher Mach numbers is of interest for future research. Experimental data and direct numerical simulation (DNS) have shown a leveling off in the decrease in shear layer growth rate with increasing Mach number. This trend is not reflected in current compressibility corrections and may be important when considering flows where compressibility effects are even more pronounced. Future work on this project also includes using detached eddy simulation (DES) to model the unsteady pressure loads on the CM surface and to better understand the turbulence in jet flows. Acknowledgments This work was supported by NASA JSC under grant NNX09AN06G. The first author would like to thank Brandon Oliver, Randy Lillard, Alan Schwing, and Darby Vicker in the Aerosciences and CFD Branch at JSC for all their help in using OVERFLOW. Also, Robert Childs at NASA Ames was very helpful with questions related to the turbulence modeling. References 1 Nichols, R. H. and Buning, P. G., User s Manual for OVERFLOW 2.1 Version 2.1t, August 2008. 2 NASA Constellation Multimedia, [online], http://www.nasa.gov/missionpages/constellation/multimedia/index.html [retreived July 17, 2009]. 3 Eggers, J. E., Velocity profiles and eddy viscosity distributions downstream of a Mach 2.22 nozzle exhausting into quiescent air, Tech. rep., NASA TN D-3601, 1966. 4 Birch, F. and Eggers, J. E., A critical review of the experimental data on turbulent shear layers, Tech. rep., NASA SP 321, 1972. 5 Lau, J. C., Measurements in Subsonic and Supersonic Free Jets Using a Laser Velocimeter, Journal of Fluid Mechanics, Vol. 93, 1979, pp. 1 27. 6 Seiner, J. M., Ponton, M. K., Jansen, B. J., and Lagen, N. T., The Effects of Temperature on Supersonic Jet Noise Emission, DGLR/AIAA 14th Aeroacoustics Conference, Aachen, Germany, 1992, AIAA Paper 92-02-046. 7 Sarkar, S., Erlebacher, G., Hussaini, M. Y., and Kreiss, H. O., The Analysis and Modeling of Dilatational Terms in Compressible Turbulence, Journal of Fluid Mechanics, Vol. 227, 1991, pp. 473 493. 8 Georgiadis, N. J., Yoder, D. A., and Engblom, W. A., Evaluation of Modified Two-Equation Turbulence Models for Jet Flow Predictions, 44th AIAA Aerospace Sciences Meeting, Reno, Nevada, 2006, AIAA Paper 2006-490. 9 Abdol-Hamid, K. S., Massey, S. J., and Elmiligui, A., Temperature Corrected Turbulence Model for High Temperature Flow, Journal of Fluids Engineering, Vol. 126, 2004, pp. 844. 10 Wishart, D. P., The Structure of a Heated Supersonic Jet Operating at Design and Off-Design Conditions, Ph.D. thesis, Department of Mechanical Engineering, Florida State University, Tallahassee, FL, May 1995. 11 Gross, N., Assessment of Turbulence Models and Corrections with Application to the Orion Launch Abort System, Master s thesis, School of Aeronautics and Astronautics, Purdue University, West Lafayette, Indiana 47907, May 2010. 12 Wernet, M., Wolter, J. D., Locke, R., Wroblewski, A., Childs, R., and Nelson, A., PIV Measurements of the CEV Hot Abort Motors Plume for CFD Validation, 48th AIAA Aerospace Sciences Meeting, Orlando, FL, 2010, AIAA Paper 2010-1031. 13 Spalart, P. R. and Allmaras, S. R., A One-Equation Turbulence Model for Aerodynamic Flows, 30th AIAA Aerospace Sciences Meeting, Reno, Nevada, 1992, AIAA Paper 92-0439. 14 Baldwin, B. S. and Barth, T. J., A One-Equation Turbulence Transport Model for High Reynolds Number Wall-Bounded Flows, Tech. rep., NASA TM 102847, 1990. 15 Wilcox, D. C., Turbulence Modeling for CFD, DCW Industries, Inc., La Cañada, CA 91011, 3rd ed., 2006. 16 Menter, F., Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications, AIAA Journal, Vol. 32, 1994, pp. 1598 1605. 17 Zeman, O., Dilatational Dissipation: The Concept and Application in Modeling Compressible Mixing Layers, Physics of Fluids, Vol. 2, 1990, pp. 178 188. 18 Suzen, Y. B. and Hoffmann, K. A., Investigation of Supersonic Jet Exhaust Flow by One- and Two-Equation Turbulence Models, 36th AIAA Aerospace Sciences Meeting, Reno, Nevada, 1998, AIAA Paper 98-0322. 9 of 21

a) b) Figure 2. Grid topology for Wishart jet. (a) Overall grid system (showing every 4th grid line), (b) detail view of nozzle lip region. 10 of 21

Figure 3. Grid topology for Glenn jet (showing every 5th grid line) along centerline showing tunnel walls (black), nozzle (red), plume region (green), and freestream (blue). Figure 4. Grid renement for Wishart underexpanded jet showing centerline Mach prole. 11 of 21

a) b) Figure 5. (a) Normalized centerline velocity and (b) normalized half radius for Eggers jet comparing turbulence models. 12 of 21

a) b) Figure 6. (a) Normalized centerline velocity and (b) normalized half radius for Eggers jet comparing compressibility corrections. 13 of 21

a) b) Figure 7. Centerline quantities for heated and perfectly expanded Wishart jet. (a) Mach number, (b) Normalized temperature. 14 of 21

Figure 8. Radial Mach number profiles for heated and perfectly expanded Wishart jet. 15 of 21

Figure 9. Mach number contours for heated and underexpanded Wishart jet. (a) CC=0, TC=0, (b) CC=1, TC=0, (c) CC=1, TC=1. 16 of 21

a) b) Figure 10. Centerline quantities for heated and underexpanded Wishart jet. (a) Mach number, (b) Normalized temperature. 17 of 21

Figure 11. Radial Mach number profiles for heated and underexpanded Wishart jet. 18 of 21

a) b) Figure 12. Normalized half radius for underexpanded jet at elevated conditions. (a) Increasing pressure ratio comparing CC. (b) Increasing temperature ratio comparing TC. 19 of 21

Figure 13. Velocity magnitude contours perpendicular to freestream for Glenn jet. 20 of 21

Figure 14. Velocity magnitude profiles along Y=0 plane for Glenn jet. 21 of 21