investigation of the mechanisms of jet-engine core noise...

23
Investigation of the mechanisms of jet-engine core noise using large-eddy simulation Jeff O’Brien * , Jeonglae Kim , and Matthias Ihme Stanford University, Stanford, California 94305 As further reductions in aircraft engine noise are realized, the relative importance of reducing engine-core noise increases. In this work, a representative engine flowpath is studied to examine the mechanisms by which direct and indirect core noise propagate through the engine and affect the far-field sound radiation. The flowpath consists of a model combustor, a single-stage turbine, a converging nozzle, a near-field jet, and far-field acoustic radiation. A combination of high-fidelity simulations and low-order semi-analytic models is used to represent the generation and propagation of disturbances through the flowpath. Particular details are provided for LES calculations of combustion chamber, nozzle exhaust flow, and jet noise radiation. A one-way coupling procedure is employed for propagating disturbances from one stage of the calculations to the next, and the results show substantial changes in the far-field sound directivity and frequency spectra due to fluctuations generated by the upstream engine core. I. Introduction Core noise is one of the important sources of noise generated by modern gas-turbine engines. As jet noise and fan noise have been progressively reduced, the relative importance of core noise has increased. Its generation is often associated with combustion and the propagation of temperature inhomogeneities through the turbine stages. 1, 2 Understanding the fundamental mechanisms of core-noise generation and propagation is an essential step toward further reducing the overall noise from gas-turbine engines. It is also important to understand how core noise interacts with the engine components, since its generation and propagation can be closely linked with thermo-acoustic instabilities in the combustor. The complex mechanisms of core noise pose significant challenges to prediction. Due to excessive compu- tational costs, modeling assumptions or theoretical tools are often employed. However, due to the subtlety of noise generation and propagation, it is desired that high-speed turbulent flows within the flowpath of gas-turbine engines are simulated using consistently high-fidelity simulation tools. * Graduate Research Assistant, Department of Mechanical Engineering. Member AIAA. Post-Doctoral Fellow, Center for Turbulence Research. Member AIAA. Assistant Professor, Department of Mechanical Engineering. Member AIAA. 1 of 23 American Institute of Aeronautics and Astronautics Downloaded by STANFORD UNIVERSITY on June 24, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-0761 54th AIAA Aerospace Sciences Meeting 4-8 January 2016, San Diego, California, USA AIAA 2016-0761 Copyright © 2015 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved. AIAA SciTech

Upload: others

Post on 02-Mar-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Investigation of the mechanisms of

jet-engine core noise using large-eddy simulation

Jeff O’Brien∗, Jeonglae Kim†, and Matthias Ihme‡

Stanford University, Stanford, California 94305

As further reductions in aircraft engine noise are realized, the relative importance of

reducing engine-core noise increases. In this work, a representative engine flowpath is

studied to examine the mechanisms by which direct and indirect core noise propagate

through the engine and affect the far-field sound radiation. The flowpath consists of a

model combustor, a single-stage turbine, a converging nozzle, a near-field jet, and far-field

acoustic radiation. A combination of high-fidelity simulations and low-order semi-analytic

models is used to represent the generation and propagation of disturbances through the

flowpath. Particular details are provided for LES calculations of combustion chamber,

nozzle exhaust flow, and jet noise radiation. A one-way coupling procedure is employed

for propagating disturbances from one stage of the calculations to the next, and the results

show substantial changes in the far-field sound directivity and frequency spectra due to

fluctuations generated by the upstream engine core.

I. Introduction

Core noise is one of the important sources of noise generated by modern gas-turbine engines. As jet

noise and fan noise have been progressively reduced, the relative importance of core noise has increased. Its

generation is often associated with combustion and the propagation of temperature inhomogeneities through

the turbine stages.1, 2 Understanding the fundamental mechanisms of core-noise generation and propagation

is an essential step toward further reducing the overall noise from gas-turbine engines. It is also important

to understand how core noise interacts with the engine components, since its generation and propagation

can be closely linked with thermo-acoustic instabilities in the combustor.

The complex mechanisms of core noise pose significant challenges to prediction. Due to excessive compu-

tational costs, modeling assumptions or theoretical tools are often employed. However, due to the subtlety

of noise generation and propagation, it is desired that high-speed turbulent flows within the flowpath of

gas-turbine engines are simulated using consistently high-fidelity simulation tools.

∗Graduate Research Assistant, Department of Mechanical Engineering. Member AIAA.†Post-Doctoral Fellow, Center for Turbulence Research. Member AIAA.‡Assistant Professor, Department of Mechanical Engineering. Member AIAA.

1 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

54th AIAA Aerospace Sciences Meeting

4-8 January 2016, San Diego, California, USA

AIAA 2016-0761

Copyright © 2015 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

AIAA SciTech

Page 2: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Large-eddy simulation (LES) has proven viable in modeling high-speed, high-temperature turbulent flows

with chemical reactions. Although its industrial application is not widespread as of yet, modern computing

power has reached a point where the complex turbulent reacting flows within realistic gas-turbine engines

can be simulated with LES in their entirety without needing to resort to low-order models. The fidelity of the

LES-based approach depends critically upon an accurate description of the complex geometry often found

in practical gas-turbine engines. Thus, LES based upon unstructured-grid technology appears essential to

satisfy such requirements and has proven suitable for massively parallel computing platforms by providing

good scalability.

In this study, we investigate fundamental core-noise mechanisms in a model gas-turbine flowpath that

contains essential components of a commercial jet engine at a high-altitude cruise condition. By applying

LES to the combustor, nozzle and jet stages of the flowpath, this study captures both the generation of direct

core-noise within the combustion chamber as well as the development of turbulent and thermal fluctuations

that lead to indirect noise downstream. Using a one-way coupling procedure, the combustor output will

inform an inflow boundary condition for a single stage, high pressure turbine. The turbine stage will be

modeled using a low-order modeling technique, and the amplification of direct noise as well as the conversion

of upstream fluctuations into indirect noise in the turbine stage. Lastly, the turbine output will be used to

form an inflow boundary condition for a simulation of a subsonic heated jet and nozzle, an environment in

which the impact of core-noise has not been studied as extensively. The thermodynamic state and modeling

approach for each stage of the flow path are shown in Figure 1.

Combustor

Compressible

Single-StageTurbine

Actuator DiskTheory

Nozzle& Jet

CompressibleLESLES

Far field

Ffowcs Williams

and Hawkings

T0 = 215K

p0 = 4.5atm

M0 = 0.06

T1 = 1010K

p1 = 4.4atm

M1 = 0.15

T2 = 895K

p2 = 1.5atm

M2 = 0.48

T3 = 805K

p3 = 1.0atm

M3 = 0.90

Figure 1. Summary of the component models and thermodynamic states for the flowpath of the representative gas-turbine engine.

II. Direct core noise generation in the combustor stage

The combustor stage is modeled using the in-house code Chris, an LES-solver that has previously been

validated in similar applications.3 Chris is a tool that solves the fully compressible Navier–Stokes equations

on unstructured grids. The code is explicit and second-order accurate in time and space on arbitrary grids.

The chemical source term is modeled by the flamelet progress variable approach4 using a three-dimensional

chemistry table based on tabulated chemistry for methane–air diffusion flames. The filtered momentum

equation is closed using the Vreman model,5 and the turbulent scalar fluxes are closed using a constant

turbulent Schmidt number assumption. The resulting set of governing equations is

2 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 3: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

∂ρ

∂t+∂ρuj∂xj

= 0 (1a)

∂ρui∂t

+∂ρuiuj∂xj

= −∂p

∂xi+

∂xj

[(µ+ µt)

(∂ui∂xj

+∂uj∂xi

−2

3δij∂uk∂xk

)−

2

3ρkδij

](1b)

∂ρE

∂t+∂ρujE

∂xj=

∂xj

[(λ

cp+

µt

Prt

)∂h

∂xj

]+

∂xj(−uj p+ uiτij) (1c)

∂ρZ

∂t+∂ρujZ

∂xj=

∂xj

[(ρD +

µt

Sct

)∂Z

∂xj

](1d)

∂ρC

∂t+∂ρujC

∂xj=

∂xj

[(ρD +

µt

Sct

)∂C

∂xj

]+ ¯ωc

(Z, Z ′′2, C

)(1e)

∂ρQ

∂t+∂ρujQ

∂xj=

∂xj

[(ρD +

µt

Sct

)∂Q

∂xj

]−

[2ρD

∂Z

∂xj

∂Z

∂xj+ ρCQZ

′′2µt

Sct∆2

](1f)

where Z is the mixture fraction, C is the progress variable, Q is the mixture fraction variance and CQ is an

empirical constant set to 40. The combustor design chosen for the flowpath is the dual-swirler gas turbine

model combustor (GTMC) configuration originally studied by Meier et al.6, 7 The combustor geometry is

represented using a block-structured hexagonal mesh of roughly 5 million control volumes. The operating

condition has been scaled up from the high-power “A” condition in Meier’s work to achieve thermodynamic

consistency with the downstream components of the flowpath as well as a specific heat release consistent

with aviation engines. In spite of this change in operating condition, the chamber dynamics are quite similar

to those observed in Meier’s high-power “A” case, as can be seen in the time-averaged flow fields shown

in Figure 2. It can be seen that the flame exhibits the same “V” configuration observed by Meier at high

power, and that the flow field is dominated by the presence of a precessing vortex core (PVC) and strong

recirculation zone in the center of the chamber.

Acoustic measurements are taken near the exit of the combustion chamber at the plane x = 31 cm. To

reduce the potential for spurious reflection, the combustor mesh is then stretched over an additional 3 chimney

diameters and characteristic non-reflecting boundary conditions are applied at the outflow boundary of the

domain. Figure 3 displays the planar-averaged spectrum taken at the combustor exhaust. The OASPL in this

plane is 155 dB, which is consistent with limited experimental data available at this location.8 Additionally,

the noise peaks at low frequencies (f < 1kHz) which is characteristic for direct combustion noise. The

spectrum exhibits several features which are indicative of the flame dynamics.

To establish the link between the flame behavior and the acoustic signature emerging from the flame, the

dynamics of the precessing vortex core are analyzed in detail. Since a connection to frequency is desired,

dynamic mode decomposition (DMD),9 is applied to examine the temporal dynamics of the flame. Specif-

ically, the sparsity promoting DMD method10 is applied to help isolate key dynamics and improve mode

3 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 4: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Figure 2. Time-averaged flow fields in GTMC combustor: a) pressure, b) axial velocity and c) temperature

Frequency [Hz]10 2 10 3 10 4 10 5

SP

L [d

B]

115

120

125

130

135

140

145

150

155

160

Figure 3. Averaged pressure spectra at combustor exit

orthogonality. Figure 4 plots the spectral coherence of the PVC’s pressure modes as a function of frequency.

The plot indicates significant mode amplitudes at 2900, 3600, and 5500 Hz. The first mode is a Helmholtz

or volumetric mode and can clearly be observed in the combustor exit spectra at this fundamental frequency

(2900 Hz) as well as its first (5800 Hz) and second harmonics (8700 Hz). The second PVC mode, at 3600

Hz, corresponds to a spinning mode in the PVC. This spinning can be observed at twice the frequency (7200

Hz) at the combustor chimney plane. While there is no peak at 7200 Hz in Figure 3 (there is a local dip in

SPL), individual microphone measurements taken in this plane exhibit strong tones at 7200 Hz as can be

seen in Figure 5. This is because phase averaging hides the presence of circumferential modes in Figure 3,

4 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 5: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Frequency [Hz]0 2000 4000 6000 8000 10000

Ene

rgy

Coh

eren

ce10 4

10 5

10 6

10 7

10 8 Pressure

Figure 4. DMD mode spectra of combustor PVC, showing mode coherence vs. frequency

Frequency [Hz]10 2 10 3 10 4 10 5

SP

L [d

B]

135

140

145

150

155

160

165

170

Figure 5. Pressure spectra taken from a single probe at the radial location r = 30 mm, x = 29 cm.

and the SPL deficit near 7200 Hz suggests that some of the available broadband energy is absorbed in the

mean flow. Lastly, the third prominent mode in the PVC is another Helmholtz mode, which can also be

observed in Figure 2. Clearly, the acoustic signature emerging from the chimney has strong links to the

in-chamber dynamics.

III. Turbine stage

The turbine stage in the flowpath consists of a single rotor and stator with a nominal pressure ratio of

2.4, allowing for a relatively large enthalpy extraction over a compact computational domain. The turbine

design is taken from Stabe et al.11 The turbine stage inflow is decomposed into three quantities:

φ(r, t) = ¯φ+ φ(r) + ψ(t)φ(r), (2)

where ¯φ is the nominal inflow operating condition of the device, φ(r) is the spatially varying, time-averaged

5 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 6: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

mean which allows the combustor’s mean flow structures to be retained, and the product ψ(t)φ(r) represents

the time and space varying fluctuations measured at the combustor exit.

For computational efficiency, the perturbations are imposed via a modal representation. Specifically,

Proper Orthogonal Decomposition (POD) is applied to the combustor outflow data, allowing for an efficient

representation of the mode shapes. Since the POD procedure can only be applied to a single quantity and

it is desired that each of the fluctuating quantities be represented by the same basis, the L2-norm of the

non-dimensional fluctuations is used:

ξ =p′

γp0+s′

cp+v′

c0(3)

where p0 represents the mean pressure, cp the specific heat of the gas, and c0 the sound speed at the combustor

exit, is chosen as the metric for the procedure, as it applies roughly equal weighting to the acoustic, vortical

and entropic perturbations. The efficiency of the POD-procedure in capturing the fluctuating fields is shown

in Figure 6 and indicates that while this metric does not perfectly capture all three fluctuating quantities

in the first few modes, it does provide an adequate representation of the fields efficiently. At this time, only

the first (largest amplitude) POD mode is imposed at the turbine inlet, and examining the effect of this

simplification is an area for further research. The fundamental mode shapes for each quantity are shown in

Figure 7.

Mode0 5 10 15 20

Mod

e A

mpl

itude

0.014

0.016

0.018

0.02

0.022

0.024

0.026

0.028

0.03

0.032

(a)

Mode0 5 10 15 20

Mod

e A

mpl

itude

0.014

0.016

0.018

0.02

0.022

0.024

0.026

0.028

0.03

0.032

(b)

Mode0 5 10 15 20

Mod

e A

mpl

itude

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

(c)

Figure 6. RMS of fluctuation amplitude for a)pressure, b) axial velocity and c) entropy vs. POD mode number.

The effect of the turbine stage is modeled using Actuator Disk Theory (ADT), a low-order analytical

method based on a linearization of conservation equations applied across a turbine stage, specifically the

conservation of momentum, energy, and entropy as well as the Kutta condition. ADT is best suited for

acoustically compact disturbances in which the characteristic length scale is much greater than the chord

lengths of the turbine blades. Because of its linearity, the method cannot take into account for the “chopping”

effect of the turbine blade-passing frequency which can map moderate frequency disturbances to higher

frequencies due to blade rotation. However, this effect is not expected to be significant at the low frequencies

6 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 7: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

r\R [-]0 0.2 0.4 0.6 0.8 1

φp

-0.22

-0.21

-0.2

-0.19

-0.18

-0.17

(a)

r/R [-]0 0.2 0.4 0.6 0.8 1

φv

-0.35

-0.3

-0.25

-0.2

-0.15

-0.1

(b)

r/R [-]0 0.2 0.4 0.6 0.8 1

φs

0.12

0.14

0.16

0.18

0.2

0.22

0.24

0.26

(c)

Figure 7. Fundamental POD-mode shapes from combustor exhaust: a) pressure, b) axial velocity and c) entropy.

associated with core noise. The method was first presented by Cumptsy and Marble,12 and has been

validated against higher fidelity methods and found to perform satisfactorily for the frequencies of interest

in this study.13, 14 The set of linearized equations is given as:

s′1 = s′2 (4a)

p′1 +w′

1

Ma1− θ′1tan(θ1) = p′2 +

w′2

Ma2− θ′2tan(θ2) (4b)

1

1 + 12(γ − 1)Ma21

(p′1 +

1

γ − 1s′1 +Ma1w

′1

)=

1

1 + 12(γ − 1)Ma22

(p′2 +

1

γ − 1s′2 +Ma2w

′2

)(4c)

θ′2 = 0 (4d)

ADT is derived from jump conditions of conserved quantities over each blade row and has been shown

to be successful at predicting the amplification of direct noise and generation of indirect core-noise at low

frequencies in multiple applications.15, 16 Figure 8 demonstrates the increase in noise (attributable to both

direct and indirect noise) across the turbine.

IV. Responses of an Mach 0.9 heated jet to the turbine-stage fluctuations

In the modeled engine flowpath, fluctuations from the turbine stage propagate downstream to an exhaust

nozzle. While propagating, some of the fluctuations decay or disperse depending on frequency, amplitude,

geometry, and so forth, and some of them convect or amplify.17 As they reach the nozzle exit, they interact

with a turbulent jet developing downstream of the nozzle.

It is well known that jet flows in general respond to external disturbances.18 Mechanisms by which

jet flows respond are complicated and sometimes not described by a simple linear theory even for small-

amplitude excitation. One of the important outcomes of such responses is a change in acoustic radiation.

7 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 8: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

r/R0 0.2 0.4 0.6 0.8 1

SP

L [d

B]

150

155

160

165

170

175

180Combustor ExitTurbine Exit

Figure 8. Pointwise OASPL increase over the turbine.

Radiated sound takes only a tiny portion of fluctuation energy of high-speed turbulent jet19 and thus, even

small-amplitude disturbances can have significant impacts on sound.

The acoustic responses of high-speed turbulent jets are often examined under conditions subject to

excitation at well-defined discrete frequencies. However, in realistic operating conditions of aircraft engine,

post-turbine fluctuations are typically broadbanded. Considering complexities of jet responses to external

excitation, turbulent jet noise subject to realistic turbine-stage fluctuations can be drastically different from

acoustics of isolated, laboratory-scale jet.

In this section, responses of a subsonic heated jet to the turbine-stage fluctuations are discussed. The

post-turbine fluctuations are imposed at the nozzle inlet as a unsteady boundary condition. The nominal

nozzle-exit condition is thermodynamically consistent with the upstream combustor and turbine stages.

The spectrum of the imposed fluctuations is broadbanded, and their magnitudes are close to the level of

fluctuations found at a realistic jet-engine’s nominal operating condition. However, the magnitudes are not

large enough to modify the mean flow and thus, the fluctuations are still considered as perturbations to the

base flow. The flow and computational conditions are first described, and the results for the baseline unforced

and forced jets are presented. Comparisons between the unforced and forced jets show that engine-core noise

has significant impacts on radiated sound of high-speed turbulent jet at high exhaust temperature.

IV.A. Simulation set-up

The high-fidelity, computational fluid dynamics code, CharLES†, is used to solve the fully compressible

Navier–Stokes equations. Ideal gas is assumed with γ = cp/cv = 1.4, where cp and cv are the specific

heats at a constant pressure and volume, respectively. Molecular viscosity is assumed to be a function of

temperature only and to follow the power-law relation, µ/µ∞ = (T/T∞)n where n = 0.76. The Prandtl

† http://www.cascadetechnologies.com/pdf/CHARLES.pdf

8 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 9: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

number is assumed constant at Pr = 0.7. The subgrid-scale model of Vreman5 is used with the model

constant c = 0.07 and a constant turbulent Prandtl number of Prt = 0.9. Turbulent jet simulations are

performed using LES under a similar condition as that of O’Brien et al.20

The governing equations are discretized in space using a cell-based finite volume formulation. Solutions

are time advanced using the standard third-order Runge–Kutta method at a constant time-step size of

∆trJ/c∞ = 0.00125, which results in the CFL number of approximately 1.0. The spatial discretization

is non-dissipative and formally second-order accurate on arbitrary unstructured grids. In addition, the

convective fluxes are combined, depending on local grid quality (for example, element skewness), with fluxes

computed by an HLLC-upwind discretization. Khalighi et al.21 provide more detailed discussions on the

spatial discretization.

High-temperature gas from the turbine stage flows into a straight pipe shown in Figure 9(a), after which

it is discharged into the ambient air at rest through an axisymmetric nozzle. The nozzle is strictly converging

and its geometry is generated to meet the ASME-standards, as can be seen in Figure 9(b). The nozzle-exit

radius is rJ = DJ/2 = 1 in. The nozzle cross-sectional area decreases over 2.5rJ in the axial direction, and

the area contraction ratio is 4.73. The straight pipe connected to the ASME-nozzle has a radius of 2.17rJ

and a length of 15rJ .

(a)

-2.0 -1.0 0.0

-4.0

-2.0

0.0

2.0

4.0

x/rJ

y/r J

(b)

Figure 9. (a) Three-dimensional geometry of the jet exhaust system. (b) Cross-section of the ASME-standard con-verging nozzle.

The nozzle-exit condition is obtained based upon the outflow condition of the turbine stage, the same as

the nominal nozzle-exit condition of the test-point number 49 of Tanna.22 The non-dimensional velocity at

the nozzle exit with respect to the ambient speed of sound is uJ/c∞ = 1.48, and the nozzle-exit temperature

relative to the ambient temperature is TJ/T∞ = 2.857. The nozzle-exit Mach number isMJ = uJ/cJ = 0.876,

9 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 10: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

and the Reynolds number based on the nozzle-exit condition is ReJ = ρJuJDJ/µJ = 2.3×105. This Reynolds

number is lower than the critical Reynolds number 4 × 105 above which low Reynolds number effects on

far-field sound become less significant.23

The physical domain extends 40rJ downstream of the nozzle exit and 40rJ in the radial direction. The

center of the nozzle exit is (x, r =√y2 + z2) = (0, 0), which is also the reference point to define far-field

locations using the distance d and the radiation angle ϕ, where ϕ = 0◦ corresponds to the downstream jet

axis, as illustrated in Figure 10.

x/rJ

y/r J

T/T∞

Figure 10. Part of computational domain on the x-y plane. The contours represent instantaneous static temperaturenormalized by the ambient temperature.

At the nozzle inlet, the total pressure of p0/p∞ = 1.621 and the total temperature of T0/T∞ = 3.229

are prescribed following the test-point number 49 of Tanna.22 At the outflow, an absorbing buffer zone is

used. The nozzle wall is modeled as a no-slip, isothermal boundary at Twall = T∞. The rest of the domain

boundaries are modeled to have the ambient total pressure and the ambient total temperature. There is no

co-flow in this flow configuration.

A base-grid having 0.8 million unstructured control volumes is generated and subsequently refined using

Adapt, the grid-adapting tool in the CharLES suite of codes. For computational accuracy and efficiency,

only hexahedral elements are used. The total number of control volumes for the adapted grid is 25.3 million.

Neither turbulent statistics nor sound prediction exhibited sensitivity upon additional mesh refinement. A

cross-section of the computational grid on the z = 0 plane is shown in Figure 11(a).

Using the same computational code CharLES, Bres et al.24 simulated a Mach 0.9 isothermal jet and its

acoustics. Their BL16M WM is compared for space–time resolution with the current simulation. It should

be noted that the nozzle geometry and nozzle-exit condition differ between Bres et al.24 (convergent–straight

nozzle and Mach 0.9 isothermal jet) and the current set-up (converging nozzle and Mach 0.9 heated jet), and

10 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 11: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

thus the comparison is qualitative. However, it can provide a useful guideline to assess whether the current

resolution is reasonable.

To investigate the impacts of turbulent boundary layer on sound radiation, Bres et al.24 selectively

refined the near-wall regions of the straight nozzle, while, in the current study, the entire region within the

nozzle is refined to resolve the fluctuations from the turbine stage (see Figure 11(b)). For the same reason,

the grid within the potential core is refined. When the nozzle interior and the potential core are not refined,

the total number of control volumes is approximately 18 million, which is comparable to Bres et al.24 In

addition, The nozzle-exit velocity of the current simulation is 1.64 times faster due to the higher temperature

ratio (TJ/T∞ = 2.857 compared to TJ/T∞ = 1.0 of Bres et al.24). As a result, the nozzle-exit boundary

layer becomes thinner for the current simulation Also, the strictly converging ASME-type nozzle suppresses

turbulent fluctuations within the incoming boundary layer, thus making the boundary layer thinner at the

nozzle exit. Overall, the resolution requirement for the current simulation appears to be higher than Bres et

al.24 This explains the choice of a smaller time-step size (∆trJ/c∞ = 0.00125 compared to ∆trJ/c∞ = 0.002

of Bres et al.24).

x/rJ

y/r J

(a)

x/rJ

y/r J

(b)

Figure 11. Cross-sections of the computational grid on the x-y plane for (a) downstream of the nozzle and (b) thenozzle interior. The dashed line in Figure 11(a) represents a part of the integral surface for the Ffowcs Williams andHawkings method.

Sound radiation at acoustic far-field locations is computed using the Ffowcs Williams and Hawkings

method.25 Additional details on the formulation and some practical guidelines for using the Ffowcs Williams

and Hawkings method are found elsewhere.26, 27 The integral surface is located within the grid-refined zones,

as illustrated in Figure 11(a). Its minimum radius is 2.9rJ near the nozzle exit and the maximum radius is

8.9rJ at the downstream end at x/rJ = 40. It was confirmed that the acoustic prediction is insensitive to

11 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 12: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

the radial location of the integral surface. Following Shur et al.,28 the end caps are used to accurately close

the Ffowcs Williams and Hawkings surface downstream. Sixteen end caps are used for 30 ≤ x/rJ ≤ 40. The

upstream end is closed by the nozzle wall and thus no end caps are required. Solutions are sampled on the

integral surface at a rate of StD ≈ 10.8. The maximum resolved frequency is estimated as StG ≈ 1.54 by

computing the grid Strouhal number29, 30

IV.B. Baseline clean-nozzle simulation

The simulation is time-averaged for 10 nominal acoustic flow-through times. Figure 12 shows the comparisons

of time-averaged axial velocity and fluctuating axial velocity rms with the particle image velocimetry (PIV)

measurement31 along the jet centerline and the nozzle lipline, respectively. The “consensus” dataset (i.e. a

weighted average of the as-measured PIV data to account for uncertainties) is compared with the current

LES prediction. Also shown in Figure 12(a) is the measurement data obtained for the Acoustic Research

Nozzle (ARN) 2. The ARN2 nozzle is a converging nozzle with the same exit diameter as the current

ASME-standard nozzle. Its contracting section is three times longer and the area contraction ratio is two

times larger than the current nozzle. Also, the nozzle internal contour is slightly different. Agreement with

both PIV-datasets is good along the jet centerline and the nozzle lipline. Flow within a strictly converging

nozzle undergoes a strong acceleration and the nozzle-exit boundary layer is likely to be relaminarized (see,

for example, Mi et al.32). This is also the case for the ARN-type nozzles,31 which can presumably explain

its slightly better agreement than the consensus dataset in Figure 12(a). As a result, velocity fluctuations at

the current nozzle exit are much lower (0.32% of uJ) compared to the experiment, as shown in Figure 12(a).

This causes the nozzle-exit boundary layer to undergo rapid transition to turbulence, substantiated by an

initial peak in u′x,rms near the nozzle exit in Figure 12(b). Artificial inflow turbulence based upon the digital

filtering technique33 was tested to increase the fluctuation levels at the nozzle exit; however, due to the strong

contraction of the current nozzle, even unrealistically strong fluctuations showed negligible improvement.

Radial profiles of streamwise velocity at several streamwise locations compare well with the consensus

PIV-measurement, as shown in Figures 13(a) and 13(b). At the nozzle exit, the time-averaged velocity has

a nearly top-hat profile. The computed 99% boundary-layer thickness is 0.0475DJ which is resolved by 20

points. The momentum thickness is 0.0024DJ where two points are used. The shape factor is computed

as 2.2, indicating that the nozzle-exit boundary layer is nominally laminar. As shown in Figure 13(b),

fluctuation levels are very low near the center of the nozzle exit and their maximum value is less than 13%

of uJ near the nozzle wall.

Figure 14(a) shows sound directivity at 72DJ from the nozzle exit. Also shown are the measurement

data of Tanna22 and Brown & Bridges.34 Agreement is good within 2 dB over the angles considered in this

12 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 13: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

0 10 20 30 400.0

0.2

0.4

0.6

0.8

1.0

1.2

0.0

0.1

0.2

0.3

0.4

x/rJ

ux/uJ

u′ x,rms/uJ

PIV31 (consensus)PIV31 (ARN2)

(a)

0 10 20 30 400.0

0.1

0.2

0.3

0.4

0.5

0.6

0.0

0.1

0.2

0.3

0.4

x/rJ

ux/uJ

u′ x,rms/uJ

PIV31 (consensus)

(b)

Figure 12. Streamwise variation of time-averaged axial velocity and fluctuating axial velocity rms along (a) the centerline(r = 0) and (b) the nozzle lipline (r/rJ = 1).

-3 -2 -1 0 1 2 30.0

0.2

0.4

0.6

0.8

1.0

y/rJ

ux/uJ

(a)

-3 -2 -1 0 1 2 30.00

0.05

0.10

0.15

0.20

y/rJ

u′ x,rms/uJ

(b)

Figure 13. Radial profiles of (a) time-averaged axial velocity and (b) fluctuating axial velocity rms at several axiallocations on the x-y plane: x/D = 0, solid; x/DJ = 4, dashed; x/DJ = 8, dashed-dot; x/DJ = 12, dashed-dot-dot;x/DJ = 16, dotted; symbols, PIV31 (consensus).

study. At sideline and upstream angles (ϕ & 90◦), the measurement of Tanna22 shows overprediction by 2

to 3 dB, which is attributed to its very large (≈ 36) area contraction ratio of the nozzle.35 Figure 14(b)

shows sound pressure levels (SPL) at the aft and sideline angles. At ϕ = 30◦, SPL is well predicted over

the computed frequency range. At the sideline angle, sound at StD & 1.0 is overpredicted by 2 to 5 dB. At

both angles, the computed sound matches well up to the estimated grid cut-off frequency StG = 1.54, shown

as a dotted line in Figure 14(b). It was confirmed that varying the radial location of Ffowcs Williams and

13 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 14: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Hawkings surface gives no improvement in Figures 14(a) and 14(b).

20 40 60 80 100 120 140

85

90

95

100

105

110

ϕ (◦)

OASPL(dB)

CurrentTanna22

Brown & Bridges34

(a)

10-2 10-1 100

90

100

110

120

130

StD

SPL(dB)

ϕ = 30◦

ϕ = 90◦

(b)

Figure 14. (a) Sound directivity and (b) sound pressure levels at ϕ = 30◦ and 90◦. Measurement is made at d/DJ = 72.

Regarding the 2 to 5 dB overprediction at StD & 1.0 at ϕ = 90◦ in Figure 14(b), it is worthwhile

mentioning the work of Viswanathan & Clark,36 who studied the effects of nozzle internal contour on radiated

sound. Their finding was that among three different converging nozzles, having the same nozzle-exit diameter

and the same operating condition, the ASME-standard nozzle (also used in this study) generates more sound.

They observed an increased SPL (≈ 3 dB) at the sideline and upstream radiation angles, especially at higher

frequencies than the spectral peaks. This trend becomes more pronounced as jet temperature increases. They

argued that different nozzle internal contours produce distinctly different boundary-layer characteristics at

the nozzle exit, thereby affecting the early development of the mixing layer and turbulence. Their Figure 12

shows similar 2 to 3 dB more sound at StD & 0.5 at the sideline angle as the current jet. Similar studies

were done by Zaman37 and Bogey & Marsden.38 Since the grid cut-off frequency is StG = 1.54, there could

be some contribution from unresolved fluctuations at StG & 1.54 on the integral surface. Nevertheless, the

overprediction seems consistent with the previous studies using the ASME-standard nozzle.

IV.C. Forced jet simulation

The baseline jet is perturbed using the fluctuations from the turbine stage. The fluctuations are prescribed

as a reference state for the inlet buffer zone having a streamwise lengh of DJ . Instantaneous temperature

contours for the baseline clean-nozzle and forced jet simulations are shown in Figures 15(a) and 15(b),

respectively. The prescribed fluctuations convect downstream through the nozzle and interact with the

turbulent jet. Jet spreading and large-scale structures do not show strong sensitivity to the upstream

14 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 15: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

fluctuations, at least qualitatively.

x/rJ

y/r J

T/T∞

(a)

x/rJ

y/r J

T/T∞

(b)

Figure 15. Instantaneous temperature contours on the x-y plane (a) without and (b) with the turbine-stage fluctuations.

Figures 16(a) through 16(c) show fluctuation amplitudes along the centerline. For the forced jet, the

maximum amplitude of the streamwise velocity fluctuation within the nozzle is 0.018uJ , corresponding to

10 m/s. The temperature fluctuation is less than 20 K and pressure fluctuation less than 165 dB. The

fluctuations do not decay significantly in the axial direction as they convect.

In Figure 17(a), streamwise velocity at r = 0 for unforced and forced jets is plotted. At the nozzle exit,

fluctuation increases from 0.32% (unforced jet) to 1.8% (forced jet) of uJ (a sufficiently large amplitude to

enter the nonlinear regime for the forced case), and time-averaged velocity increases by 0.024uJ . Due to

the increased level of upstream fluctuation, forced jet decays faster and its centerline fluctuation saturates

slightly earlier than that of unforced jet. At the nozzle lipline, the initial u′x,rms peak discussed in Figure 12(b)

persists for the forced jet, though its amplitude is slightly reduced due to the increased upstream fluctuation.

This implies that the nozzle-exit boundary layer is still nominally laminar.

Figure 18(a) shows the comparisons of radial profiles of streamwise velocity at several axial locations

for unforced and forced jets. At the nozzle exit, the impacts of the upstream fluctuations are significant.

The 99% boundary-layer thickness of the forced jet is 0.305DJ , nearly 6.4 times larger than that of the

unforced jet. In addition, the momentum thickness increases four times due to the upstream fluctuations.

Boundary-layer shape factors for unforced and forced jets are 2.2 and 1.28, respectively. However, the shape

factor of the forced jet does not necessarily indicate that its nozzle-exit boundary layer is fully turbulent.

Rather, this appears to be caused by large-scale thermal fluctuations that modify the velocity profiles.

A comparison of the sound directivity at d/DJ = 72 is shown in Figure 19(a) for unforced and forced jets.

15 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 16: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

-15 -10 -5 0 5 100.00

0.02

0.04

0.06

x/rJ

u′ x,rms/uJ

Unforced jetForced jet

(a)

-15 -10 -5 0 5 100

10

20

30

40

50

60

70

x/rJ

T′(K

)

Unforced jetForced jet

(b)

-15 -10 -5 0 5 10120

130

140

150

160

170

x/rJ

p′(dB)

Unforced jetForced jet

(c)

Figure 16. Centerline fluctuation amplitudes of (a) streamwise velocity, (b) temperature, and (c) pressure.

Depending on the radiation angle ϕ, the incoming turbine-stage fluctuations have two distinct effects on the

radiated sound. At lower radiation angles, OASPL is decreased with its maximum reduction of 1.2 dB at

ϕ = 50◦. In the sideline and forward directions, sound is consistently amplified. The maximum amplification

is 3.1 dB at ϕ = 85◦.

In Figure 19(b), frequency spectrum at ϕ = 30◦ is plotted. At lower frequencies (StD . 0.05 or 500 Hz), a

consistent increase by ≈ 5 dB is observed. This is often associated with the direct contribution of engine-core

noise. Pressure fluctuations produced by combustion processes are characterized by frequencies at O(102) Hz.

Their wavelengths are typically much longer than the geometric length scales of engine components. Some

16 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 17: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

0 10 20 30 400.0

0.2

0.4

0.6

0.8

1.0

1.2

0.0

0.1

0.2

0.3

0.4

x/rJ

ux/uJ

u′ x,rms/uJ

Unforced jetForced jet

(a)

0 10 20 30 400.0

0.1

0.2

0.3

0.4

0.5

0.6

0.0

0.1

0.2

0.3

0.4

x/rJ

ux/uJ

u′ x,rms/uJ

Unforced jetForced jet

(b)

Figure 17. Streamwise variation of time-averaged axial velocity and fluctuating axial velocity rms for unforced andforced jets along (a) the centerline (r = 0) and (b) the nozzle lipline (r/rJ = 1).

-3 -2 -1 0 1 2 30.0

0.2

0.4

0.6

0.8

1.0

y/rJ

ux/uJ

(a)

-3 -2 -1 0 1 2 30.00

0.05

0.10

0.15

0.20

y/rJ

u′ x,rms/uJ

(b)

Figure 18. Radial profiles of (a) time-averaged axial velocity and (b) fluctuating axial velocity rms at several axiallocations on the x-y plane: x/D = 0, solid; x/DJ = 4, dashed; x/DJ = 8, dashed-dot; x/DJ = 12, dashed-dot-dot;x/DJ = 16, dotted. Lines with symbols represent forced-jet results.

of the pressure fluctuations are able to directly propagate through the engine flowpath through the exhaust

nozzle and acoustic far field. In addition, the indirect process generates additional pressure fluctuations at

the combustor nozzle exit and turbine stages, some of which can propagate to the acoustic far field.1 The

observed noise amplification at StD . 0.05 shows the core-noise propagation unambiguously contributes to

far-field sound of jet engines.

At frequencies higher than StD ≈ 0.05 in Figure 19(b), radiated sound is suppressed. The noise reduction

17 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 18: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

is approximately uniform over frequencies as ≈ 5 dB. Also, several predominant tones are found at discrete

frequencies. This observation is consistent with the broadband jet-noise reduction.39, 40 When jet flows

are perturbed using acoustic excitation, radiated sound is often reduced over a range of frequencies where

jet noise is dominant. The acoustic excitation triggers the most unstable (fundamental) shear-layer mode

at 0.01 ≤ Stθ = f0θ/uJ ≤ 0.02. As its amplitude grows and saturates downstream, energy is gradually

transferred to subharmonic motions. This is demonstrated as pronounced peaks at subharmonic frequencies

of f0 and successive vortex pairing in the jet shear layer. Hence, shear-layer dynamics becomes more regular

and broadband turbulence is suppressed, which leads to the broadband reduction of radiated noise. Due to

the tonal amplification at f0 and its subharmonics, far-field sound shows strong peaks at the corresponding

frequencies.

20 40 60 80 100 120 140

90

95

100

105

110

ϕ (◦)

OASPL(dB)

Unforced jetForced jet

(a)

10-2 10-1 100

10-4 10-3 10-2

80

90

100

110

120

130

StD

SPL(dB)

Stθ

Unforced jetForced jet

f0

f0/2

(b)

Figure 19. (a) Sound directivity and (b) sound pressure levels at ϕ = 30◦ for unforced and forced jets. Measurementis made at d/DJ = 72.

Figure 19(b) shows that the broadband jet-noise radiation is suppressed at StD & 0.05 and tonal am-

plification is observed at the shear-layer instability frequencies denoted by f0 and subharmonic f0/2. The

excitation of the most unstable shear-layer mode and its subharmonics is evidenced by examining near-field

pressure spectra. As illustrated in Figure 20(a), pressure amplitudes at several axial locations along the

nozzle lipline are computed. Figure 20(b) shows pressure spectra for unforced and forced jets at x/rJ = 3.

For the unforced jet, spectra are broadbanded and distinct peaks are not observed. In contrast, for the

forced jet, two strong peaks exist near f0, which correspond to StD ≈ 1.0 (Stθ ≈ 0.01). The stronger peak

at StD = 0.46 (Stθ = 0.0046) represents the subharmonic mode. At x/rJ = 7, the subharmonic becomes

predominant and the fundamental mode decays to the level of broadband turbulent fluctuations, as can be

seen in Figure 20(c). In Figure 20(d), farther downstream at x/rJ = 11, fluctuation peaks at StD = 0.28.

18 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 19: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Spectrum becomes more broadbanded and similar with that of the unforced jet.

x/rJ

y/r J

T/T∞

(a)

10-2 10-1 100 101

10-4 10-3 10-2

10-4

10-3

10-2

StD

PSD

Stθ

Unforced jetForced jet

f0f0/2

(b)

10-2 10-1 100 101

10-4 10-3 10-2

10-4

10-3

10-2

StD

PSD

Stθ

Unforced jetForced jet f0/2

(c)

10-2 10-1 100 101

10-4 10-3 10-2

10-4

10-3

10-2

StD

PSD

Stθ

Unforced jetForced jet

(d)

Figure 20. (a) Pressure-measured locations denoted by circles. (b-d) Lipline pressure spectra at x/rJ = 3, 7, and 11,respectively, for unforced and forced jets.

V. Summary and future work

A hybrid modeling approach to predict engine core noise from a modeled gas-turbine engine and assess

its receptivity to nozzle upstream fluctuations is proposed. The modeled core-noise system consists of a

combustor, single-stage turbine, converging nozzle, and free-field radiation to the acoustic far field. The

computational strategy for the generation and propagation of turbulent fluctuations from the combustor to

the nozzle exhaust is developed.

19 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 20: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Compressible reacting LES is performed for flows within the modeled gas-turbine combustor. Fluctuation

data are recorded at the combustor exit. A modal analysis of the combustor’s in-chamber behavior is

performed, and a clear link between the precessing vortex core and combustor exit acoustics is established.

POD is then applied to decompose the signals coming from the combustor for an efficient downstream

coupling with the turbine stage. Effects of the turbine stage on the fluctuating fields, in turn, are simulated

using the semi-analytic ADT technique to estimate the fluctuations that would be seen at the exhaust-nozzle

entrance. These fluctuations are then used to forced a subsonic heated jet and to assess the changes induced

by the upstream perturbations on acoustic radiation.

High-fidelity simulations of the high-temperature jet exhaust flow are conducted using the combined

LES and acoustic analogy based upon the Ffowcs Williams and Hawkings formulation. Good agreement is

obtained for jet turbulence and sound radiation. Some discrepancies remain for the upstream propagating

sound at higher frequencies.

The baseline jet is then perturbed by fluctuations generated by the POD–ADT technique modeling the

effects of the turbine stage. Fluctuation levels at the nozzle exit are comparable with the experiment and the

forced jet decays slightly faster than the baseline jet. Far-field sound is found to decrease slightly over shallow

downstream angles (≈ 1 dB) while the upstream directivity is amplified (≈ 3 dB). The sound spectra show

that the prescribed forcing has two main effects on downstream radiation: low frequency sound is amplified

while high frequency sound is reduced and exhibits clear tones. The increase in low frequency noise suggests

that the upstream combustion noise passes through the nozzle and jet largely unimpeded and radiates at

all angles. The high frequency effects are likely attributable to an increase in vortex pairing at the nozzle

lip resulting in decreased turbulence in the shear layer and pronounced tones indicative of the characteristic

eddy size.

Future work will focus on more detailed analysis of the effects of upstream perturbations on downstream

noise, including attempts to assess the relative importance of direct and indirect core noise. More sophisti-

cated coupling techniques and a higher fidelity model for the turbine stage will also be pursued.

Acknowledgments

The authors acknowledge the following award for providing computing and visualization resources that

have contributed to the research results reported within this paper: MRI-R2: Acquisition of a Hybrid

CPU/GPU and Visualization Cluster for Multidisciplinary Studies in Transport Physics with Uncertainty

Quantification (http://www.nsf.gov/awardsearch/showAward.do?AwardNumber=0960306). This award is

funded under the American Recovery and Reinvestment Act of 2009 (Public Law 111-5). Additional comput-

ing resources are provided by the Argonne National Laboratory through the ASCR Leadership Computing

20 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 21: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

Challenge.

The first author acknowledges the Stanford Graduate Fellowship program for continued support of this

work. The authors are grateful to Prof. Sanjiva Lele for useful discussions. The authors also thank Dr. James

Bridges at the NASA Glenn Research Center for sharing the PIV and acoustic measurement data for vali-

dation.

References

1Strahle, W. C., “Combustion noise,” Prog. Energy Combust. Sci., Vol. 4, No. 3, 1978, pp. 157–176.

2Hubbard, H. H., “Aeroacoustics of flight vehicles: theory and practice. Volume 1. Noise sources,” Tech. rep., DTIC

Document, 1991.

3See, Y. C. and Ihme, M., “Large eddy simulation of a partially-premixed gas turbine model combustor,” Proc. Combust.

Inst., Vol. 35 (in press.), 2014.

4Pierce, C. and Moin, P., “Progress-variable approach for large-eddy simulation of non-premixed turbulent combustion,”

J. Fluid Mech., Vol. 504, 2004, pp. 73–97.

5Vreman, A. W., “An eddy-viscosity subgrid-scale model for turbulent shear flow: Algebraic theory and applications,”

Phys. Fluids, Vol. 16, No. 10, 2004, pp. 3670–3681.

6Weigand, P., Meier, W., Duan, X. R., Stricker, W., and Aigner, M., “Investigations of swirl flames in a gas turbine

model combustor: I. Flow field, structures, temperature, and species distributions,” Comb. and Flame, Vol. 144, No. 1, 2006,

pp. 205–224.

7Meier, W., Duan, X., and Weigand, P., “Investigations of swirl flames in a gas turbine model combustor: II. Turbulence-

chemistry interactions,” Comb. and Flame, Vol. 144, No. 1, 2006, pp. 225–236.

8Bake, F., Kings, N., Fischer, A., and Rohle, I., “Experimental investigatino of the entropy noise mechanism in aero-

engines,” Int. J. Aeroacoustics.

9Schmid, P., “Dynamic mode decomposition of numerical and experimental data,” J. Fluid Mech., Vol. 656, 2010, pp. 5–28.

10Jovanovic, M., Schmid, P., and Nichols, J., “Sparsity-promoting dynamic mode decomposition,” Phys. of Fluids, Vol. 26,

2014, pp. 1–22.

11Stabe, R., Whitney, W., and Mofitt, T., “Performance of a high-work low aspect ratio turbine tested with a realistic inlet

radial temperature profile,” 20th AIAA/SAE/ASME Joint Propulsion Conference, Cincinnati, Ohio, June 11–13 1978, AIAA

Paper No. AIAA-84-1161.

12Cumpsty, N. and Marble, F., “The interaction of entropy fluctuations with turbine blade rows; a mechanism of turbojet

noise,” Proceedings of the Royal Society of London A, Vol. 357, 1977, pp. 323–344.

13Mishra, A. and Bodony, D., “Evaluation of actuator disk theory for predicting indirect combustion noise,” Journal of

Sound and Vibration, Vol. 332, 2013, pp. 821–838.

14Duran, I., Leyko, M., Moreau, S., Nicoud, F., and Poinsot, T., “Computing combustion noise by combining Large Eddy

Simulations with analytical models for the propagation of waves through turbine blades,” Comptes Rendus Mecanique, Vol. 341,

No. 1, 2013, pp. 131–140.

15Leyko, M., Moreau, S., Nicoud, F., and Poinsot, T., “Wave transmission and generation in turbine stages in a combustion

noise framework,” 16th AIAA/CEAS Aeroacoustics conference, Stockholm, Sweden, June 7–9 2010, AIAA Paper No. AIAA-

2010-4032.

21 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 22: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

16Duran, I. and Moreau, S., “Study of the attenuation of waves propagating through fixed and rotating turbine blades,”

18th AIAA/CEAS Aeroacoustics conference, Colorado Springs, Colorado, June 4-6 2012, AIAA Paper No. AIAA-2012-2133.

17Dowling, A. P. and Mahmoudi, Y., “Combustion noise,” Proc. Combust. Inst., Vol. 35, No. 1, 2015, pp. 65–100.

18Crow, S. C. and Champagne, F. H., “Orderly structure in jet turbulence,” J. Fluid Mech., Vol. 48, 1971, pp. 547–591.

19Crighton, D. G., “Basic principles of aerodynamic noise generation,” Prog. Aerosp. Sci., Vol. 16, No. 1, 1975, pp. 31–96.

20O’Brien, J., Kim, J., and Ihme, M., “Integrated analysis of jet-engine core noise using a hybrid modeling approach,” 21st

AIAA/CEAS Aeroacoustics Conference (36th AIAA Aeroacoustics Conference), Dallas, Texas, June 22–26 2015, AIAA Paper

No. 2015–2821.

21Khalighi, Y., Nichols, J. W., Lele, S. K., Ham, F., and Moin, P., “Unstructured large eddy simulation for prediction of noise

issued from turbulent jets in various configurations,” 17th AIAA/CEAS Aeroacoustics Conference (32nd AIAA Aeroacoustics

Conference), Portland, Oregon, June 5–8 2011, AIAA Paper No. 2011–2886.

22Tanna, H. K., “An experimental study of jet noise Part I: Turbulent mixing noise,” J. Sound Vib., Vol. 50, No. 3, 1977,

pp. 405–428.

23Viswanathan, K., “Aeroacoustics of hot jets,” J. Fluid Mech., Vol. 516, 2004, pp. 39–82.

24Bres, G. A., Jordan, P., Colonius, T., Le Rallic, M., Jaunet, V., and Lele, S. K., “Large eddy simulation of a Mach 0.9

turbulent jet,” Proceedings of the Summer Program, Center for Turbulence Research, Stanford University/NASA Ames, 2014.

25Ffowcs Williams, J. and Hawkings, D. L., “Sound generated by turbulence and surfaces in arbitrary motion,” Philos.

Trans. R. Soc. London, Ser. A, Vol. 264, No. 1151, 1969, pp. 321–342.

26Bres, G. A., Nichols, J. W., Lele, S. K., and Ham, F., “Towards best practices for jet noise predictions with unstructured

large eddy simulations,” 42nd AIAA Fluid Dynamics Conference and Exhibit, New Orleans, Louisiana, June 25–28 2012, AIAA

Paper No. 2012–2965.

27Mendez, S., Shoeybi, M., Lele, S. K., and Moin, P., “On the use of the Ffowcs Williams-Hawkings equation to predict

far-field jet noise from large-eddy simulations,” Int. J. Aeroacoust., Vol. 12, No. 1-2, 2013, pp. 1–20.

28Shur, M. L., Spalart, P. R., and Strelets, M. K., “Noise prediction for increasingly complex jets. Part I: Methods and

tests,” Int. J. Aeroacoust., Vol. 4, No. 3, 2005, pp. 213–246.

29Uzun, A., Lyrintzis, A. S., and Blaisdell, G. A., “Coupling of integral acoustics methods with LES for jet noise prediction,”

Int. J. Aeroacoust., Vol. 3, No. 4, 2004, pp. 297–346.

30Bodony, D. J. and Lele, S. K., “On using large-eddy simulation for the prediction of noise from cold and heated turbulent

jets,” Phys. Fluids, Vol. 17:8, 2005, pp. 085103.

31Bridges, J. and Wernet, M. P., “The NASA subsonic jet particle image velocimetry (PIV) dataset,” NASA/TM–2011-

216807, National Aeronautics and Space Administration Glenn Research Center, Cleveland, Ohio, 2011.

32Mi, J., Nobes, D. S., and Nathan, G. J., “Influence of jet exit conditions on the passive scalar field of an axisymmetric

free jet,” J. Fluid Mech., Vol. 432, 2001, pp. 91–125.

33Klein, M., Sadiki, A., and Janicka, J., “A digital filter based generation of inflow data for spatially developing direct

numerical or large eddy simulations,” J. Comput. Phys., Vol. 186, No. 2, 2003, pp. 652–665.

34Brown, C. and Bridges, J., “Small hot jet acoustic rig validation,” NASA/TM–2006-214234, National Aeronautics and

Space Administration Glenn Research Center, Cleveland, Ohio, 2006.

35Harper-Bourne, M., “Jet noise measurements: Past and present,” Int. J. Aeroacoust., Vol. 9, No. 4-5, 2010, pp. 559–588.

36Viswanathan, K. and Clark, L. T., “Effect of nozzle internal contour on jet aeroacoustics,” Int. J. Aeroacoust., Vol. 3,

No. 2, 2004, pp. 103–135.

22 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761

Page 23: Investigation of the mechanisms of jet-engine core noise ...web.stanford.edu/group/ihmegroup/cgi-bin/MatthiasI... · Investigation of the mechanisms of jet-engine core noise using

37Zaman, K. B. M. Q., “Effect of initial boundary-layer state on subsonic jet noise,” AIAA J., Vol. 50, No. 8, 2012,

pp. 1784–1795.

38Bogey, C. and Marsden, O., “Numerical modelling of jets exiting from the ASME and conical nozzles,” 53rd AIAA

Aerospace Sciences Meeting, Kissimmee, Florida, January 5–9 2015, AIAA Paper No. 2015–0510.

39Kibens, V., “Discrete noise spectrum generated by acoustically excited jet,” AIAA Journal , Vol. 18, No. 4, 1980, pp. 434–

441.

40Hussain, A. K. M. F. and Hasan, M., “Turbulence suppression in free turbulent shear flows under controlled excitation.

Part 2. Jet-noise reduction.” Journal of Fluid Mechanics, Vol. 150, 1985, pp. 159–168.

23 of 23

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

June

24,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

0761