Applsci 12 03356 v2
Applsci 12 03356 v2
Applsci 12 03356 v2
sciences
Article
URANS Analysis of a Launch Vehicle
Aero-Acoustic Environment
Mara S. Escartí-Guillem 1,2 , Luis M. García-Raffi 1 and Sergio Hoyas 1, *
Abstract: Predicting and mitigating acoustic levels become critical because of the harsh acoustic
environment during space vehicle lift-off. This paper aimed to study the aero-acoustic environment
during a rocket lift-off. The sound propagation within a launch event was studied using dedicated
computational fluid dynamics (CFD). The resolution of all the phenomena that occur is unfeasible.
We discuss the turbulence simplification and propose a feasible simulation through an unsteady
Reynolds-averaged Navier–Stokes (URANS) model. The results were validated with experimental
data showing a good correlation near the fairing surface and an improvable accuracy in the far field.
To assess noise generation, the main shock waves were identified, and the evolution of the generated
sound pressure was assessed. Moreover, vertical directivity was revealed by data analysis of the
pressure field surrounding the fairing.
Citation: Escartí-Guillem, M.S.; 1. Introduction
García-Raffi, L.M.; Hoyas, S. URANS
The acoustic conditions during spacecraft launch are highly complex. The exhaust
Analysis of a Launch Vehicle
gas jet of the rocket engine and its collision with the launchpad produce severe acoustic
Aero-Acoustic Environment. Appl.
waves, which are the primary sound sources during lift-off [1,2]. This acoustic environment
Sci. 2022, 12, 3356. https://doi.org/
causes a severe acoustic and vibration load on the payload, endangering the mission and
10.3390/app12073356
potentially resulting in financial losses [3]. Hence, predicting and mitigating acoustic levels
Academic Editor: Manuel Armada during lift-off are critical in the launch pad and launch vehicle design since this enhances
the reliability of the launcher while also increasing payload comfort.
Received: 16 February 2022
Accepted: 18 March 2022
This work is part of a project, conducted under a European Space Agency (ESA)
Published: 25 March 2022
program, to develop a sound mitigation system applicable to the launch facility of the VEGA
rocket. The first step towards any mitigation process is the prediction of noise sources.
Publisher’s Note: MDPI stays neutral The global study of sound generation and transmission up to the payload entails several
with regard to jurisdictional claims in
problems of different natures, which are summarized in Figure 1a. First, sound generation
published maps and institutional affil-
is mainly due to the exhaust gases’ jet with the blast wave generated during ignition.
iations.
As shown in Figure 1b, the noise generated by the jet comprises three distinct sound
sources: turbulent mixing noise (TMN), shock-associated noise (SAN), and impinging
tones. The TMN is generated by the turbulent mixing of large-scale or small-scale eddies at
Copyright: © 2022 by the authors.
the mixing layer [2,4]. Large-scale vortex noise dominates at supersonic speeds, resulting
Licensee MDPI, Basel, Switzerland. in low-frequency noise with high downstream directivity, as the blue arrows represent
This article is an open access article in Figure 1b [5]. The SAN in impinging jets is produced by the interaction of turbulent
distributed under the terms and structures with shock waves, and it is called broadband shock-associated noise (BBSAN).
conditions of the Creative Commons The BBSAN is broadband and primarily directed upstream, as the green arrows represent,
Attribution (CC BY) license (https:// where the launch vehicle is located. Finally, the impinging tones appear when the jet
creativecommons.org/licenses/by/ collides and are directed upstream [2]. Second, there is an air transmission of these noise
4.0/). sources, which interact with the launchpad and are redirected to the vehicle. Thirdly, blast
wave noise is also transmitted through the launcher structure up to the fairing and payload
base. Finally, the structural vibration of the fairing interacts with the airborne noise in a
vibroacoustic problem.
Expansion waves
Jet boundary
Mach disk
Detached
shock wave
Curvature
shock wave
(a) (b)
Figure 1. Sound generation and transmission during a rocket launch. (a) Scheme of noise transmission
paths. (b) Scheme of noise sources due to the rocket plume.
This work aimed to analyze the sound sources and their propagation during the lift-off
stage. The approach here was to use computational fluid dynamics (CFD) to resolve the
aero-acoustic problem that includes the study of the engine exhaust jet and its interaction
with the launch pad. The simulation of all the physical phenomena that occur during
the launch of rockets, such as exhaust gases’ combustion, supersonic flow compressibil-
ity, and lift of the launch vehicle, is unfeasible. Solving the generated turbulence with a
sufficient resolution requires a direct numerical simulation in a vast domain [6]. There is
currently not enough computational power to achieve this. This lack of resources forces
making simplifications, as done in the literature [7,8]. Within these simplifications, it is
fundamentally under discussion up to what level turbulence should be modeled since it
has great importance in flow behavior and noise generation. In the context of supersonic
jets, large eddy simulation (LES) is the standard because it is a very accurate model, as
demonstrated against experimental data [4,9,10]. However, the computational cost of
reliably simulating full-scale spacecraft launch environments is still very high. Therefore,
we propose the simulation of this environment using the unsteady Reynolds-averaged
Navier–Stokes (URANS) model, which provides a level of simplification that is computa-
tionally feasible. The URANS model has been previously used and validated to predict
shock waves due to impinging supersonic jets as in [11] or for a VEGA fairing in transonic
flight in [12]. Furthermore, we attempted to demonstrate that these simplifications are
correct in the launchpad environment by validating with experimental data. In the previ-
ous article [13], we presented the methodology, and in this one, we tested the hypotheses
through validation and deepened the analysis of the results.
This work is organized as follows: Section 2 describes the methodology employed
during the numerical prediction modeling. Section 3 presents the results and the validation.
Section 4 summarizes the conclusions of the work.
2. Methods
The tool used for the simulations was the software OpenFOAM v6. All the details of
the numerical model can be found in the previous article [13].
Appl. Sci. 2022, 12, 3356 3 of 9
∂Uj ∂U ∂P ∂
ρ + ρUj i = − + (2µsij − ρu0j ui0 ) (2)
∂t ∂x j ∂xi ∂x j
In this work, we used the URANS equations since we retained the transient term,
∂Uj /∂t, during the computation so that the averaged components were still a function of
time. The URANS equations are the usual RANS equations as Equation (2).
The acoustic field that the launch vehicle must withstand is due primarily to the fluc-
tuating component of the turbulence of the exhaust gas mixture layer. In the acoustic field,
sound pressure (Ps ) generated by a sound wave is understood as the pressure deviation
from the ambient atmospheric pressure (P0 ). Thus, the total pressure (P) is the sum of
these two components. The nature of turbulent flows is chaotic, and therefore, they were
analyzed in statistical terms. The Reynolds decomposition states that any flow magnitude
is decomposed into a mean (Pm ) and a fluctuating part (Pf ), where Pm can be seen as an
ensemble average of the pressure measured in an infinite number of experiments. The fluc-
tuating part Pf can only be computed through expensive algorithms such as DNS. Here, the
turbulence was modeled by a URANS model. This model is based on the characterization of
the transport and dissipation of energy through an effective viscosity called eddy viscosity
introduced by the Boussinesq approximation [14]. When using the URANS turbulence
model, the pressure is simply:
P = P0 + Pm + Pf −
→ PURANS = P0 + Pm (3)
but the numerical method, through the two-equation eddy viscosity model k − ω SST [15],
takes into account all fluctuations at non-resolved scales. This model was chosen because it
is a two-layer model based on the Wilcox k − ω model in the inner parts of the boundary
layer and switches to the k − e model in the free-stream region. However, it is not possible
to estimate Pf nor its standard deviation. Therefore, Pm is a smooth function where the
high-frequency noise due to turbulent fluctuations is lost. However, due to the transient
nature of the model, information on pressure variations caused by shock waves and the
average of turbulence can be collected.
Appl. Sci. 2022, 12, 3356 4 of 9
Figure 2. Mesh details. On the left picture, mesh domain cross-section representing each of the cells.
On the right, representation of the mesh classified by the size of the cells.
the mean pressure obtained by the simulation was not within the experimentally acceptable
range. However, the experimentally acceptable range was within the mean minus the
simulation’s standard error. In this region, the performance of the CFD model was not ideal.
Factors such as the simplification of the exhaust plume input data or the movement of the
rocket led to different sources of error in both locations. However, the parameters obtained
at the plume development region and the flow behavior agreed with the literature [13],
and the results at the fairing presented a better accuracy.
1 Simulation
Simulation
0.8 Experimental Experimental
0.8
0.6 0.6
0.4 0.4
0.2
0.2
0
0
2 4 6 8 0 5 10 15 20 25
Microphone Microphone
(a) (b)
Figure 3. Time-averaged pressure validation against experimental measurements along (a) the VEGA
fairing and (b) one of the masts. Results are normalized.
Figure 4. VEGA lift-off environment shock waves’ identification. Normalized instantaneous pressure
contour at (a) t = 0.0064 s and (b) t = 0.02 s. (c) Schlieren shock waves’ identification.
Figure 4b at t = 0.02 s depicts the evolution of the blast wave due to the rocket exhaust
gases. There is a green wavefront that corresponded to the jet core boundary and a second
Appl. Sci. 2022, 12, 3356 6 of 9
one that emerged from the nozzle and moved downstream while being disrupted by
launchpad elements. The evolution of this second wavefront can be seen in the top images
of Figure 5, between 0.0267 s and 0.0377 s. It was caused by the initial exit of the gases and
their impact on the deflector. As it moved upstream, it was disrupted by the elements of the
structure. Between 0.04 s and 0.16 s, the blast wave continued to advance upstream. At the
same time, the waves generated by the interaction with the support structure separated
from the wave generated at the output of the jet. Due to the faster sound propagation
speed, the wavefront eventually overtook the fluid front and was attenuated. At the final
time step, the maximum pressure values were found in the jet development region and the
deflector due to the supersonic velocity and shock waves that appeared. On this basis, it
can be concluded that the maximum pressure levels were produced by the initial normal
shock wave and its subsequent impact with the deflector. The acoustic waves propagated
upstream of the vehicle due to this impact and the effect of the flow movement.
-∞ 0
Figure 6. Pressure temporal evolution along the fairing at (a) an azimuth angle of 315º and different
heights, (b) at a constant height of 39.7 m with different azimuth angles, and (c) at 42.7 m and different
azimuth angles.
The discrepancy in the pressure between points at the same height was small, as seen
in Figure 6b,c. However, there was an increase in this difference with the increase in height.
The maximum pressure variation between points on the same plane was about 1/10 of that
for the points at different heights. This was mainly because the fluid front reached each
plane at a different time. It should also be noted that the region at the lower part of the
fairing supported bigger pressure values than the upper region.
As the pressure gradient can be used to assess the direction of pressure waves, it
was studied from the simulation results. Figure 7a depicts the pressure gradient in each
direction, its magnitude, and the pressure gradient streamlines. The pressure gradient at
the X and Y coordinates was lower than the axial direction. The streamlines show that the
pressure gradient moved axially and converged at the tip of the payload fairing. As seen in
Figure 7b, the Z negative component was the main contributor to the pressure magnitude
at the tip. This negative component created a gradient pressure force that attracted the
flow. This gradient decreased as the pressure wave weakened, with the maximum values
occurring around t = 0.1 s. With this analysis, it is clear that the pressure field had a vertical
direction due to the movement of the exhaust gases.
DIFFUSE ACOUSTIC LOAD
o The streamlines show that the pressure gradient 𝑑𝑃! 𝑑𝑃! 𝑑𝑃! Pressure Gradient Magnitude
moves along the axial direction and converge at 𝑑𝑥 𝑑𝑦 𝑑𝑧
the tip of the payload faring.
o At the tip of the payload faring the main
contributors is the Z negative component.
oThe variations are up to 2 orders of magnitude
t = 0.126 s
(a) 13 (b)
Figure 7. Pressure gradient magnitude and components. (a) Graphical representation at t = 0.126 s.
(b) Plot over the fairing axial plane.
4. Conclusions
In the context of designing a noise-reduction system to improve sound attenuation
during space vehicles’ launch, a numerical modeling framework was defined to predict
Appl. Sci. 2022, 12, 3356 8 of 9
acoustic loads. CFD was used to assess the noise generation and propagation mechanisms
in the low-frequency regime. The reliable analysis of this type of environment involves
considering many variables and physical phenomena such as the combustion of gases,
the compressibility of the supersonic flow, and the lift of the launch vehicle. In addition,
if we want to solve the generated turbulence and with a sufficient resolution from the
acoustic point of view, the simulation of a rocket launch requires an extremely high compu-
tational and engineering cost that is out of industrial reach. In this work, the simulation
was simplified to model this complex environment from the engineering point of view.
The main requirements were to minimize the computational time while modeling the
phenomena accurately. Thus, it was evaluated whether URANS-type modeling offered
sufficient accuracy at a low computational cost.
A limitation of the work was that there were no experimental data close to the rocket
engine. Therefore, data on the far-field and fairing surface were used. The validation
showed a better data correlation at the fairing surface than in the masts surrounding the
launch pad. The apparent lack of correlation could be attributed to different sources of error
as the simplification of the launch pad structures, of the exhaust gases thermodynamics,
or URANS modeling. However, there was a bigger refinement of the cells at the fairing
rather than at the far-field points. Hence, this was reflected in the better prediction of the
results. The primary refinement was in the region where the jet develops. Upstream, the
cells’ size increased to reduce the computational cost. The approximate computational time
to simulate 0.1 “real-life” s of the 106 element mesh with 128 processors was 33 d, which is a
substantial amount. Moreover, a collateral issue was the large amount of data generated for
each step. The greater the mesh, the greater the amount of data was. Further work should
address the effect of the mesh size in this region on the accuracy and computational cost.
The results found that the sound pressure maxima appeared due to the normal shock
wave generated by the jet and its impingement with the deflector. Moreover, the shock
waves present were visualized by numerical Schlieren pictures. Furthermore, the vertical
directivity of the pressure over the fairing and the propagation of upstream sound waves
were observed. These findings showed that the diffuse field did not represent the acoustic
field that excited the fairing.
Finally, these results provided insight into the phenomena during a rocket lift-off.
The knowledge regarding the noise propagation paths will help design the noise reduction
systems and define their positioning on the launch pad.
Author Contributions: Conceptualization, S.H. and L.M.G.-R.; methodology, S.H., M.S.E.-G. and
L.M.G.-R.; software, M.S.E.-G.; validation, S.H., M.S.E.-G. and L.M.G.-R.; formal analysis, M.S.E.-G.;
resources, S.H.; writing—original draft preparation, M.S.E.-G.; writing—review and editing, S.H. and
L.M.G.-R.; supervision, S.H. and L.M.G.-R.; funding acquisition, S.H., M.S.E.-G. and L.M.G.-R. All
authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the European Space Agency of Project REDLAUCH: Launch
Sound Level Reduction under contract 4000126316/19/NL/LvH. The work was supported by the
MICINN (grants: DIN2019-010877 and RTI2018-102256-B-100) and by the Barcelona Supercomputing
Center under Project IM-2021-2-0017 Rocket launch aeroacoustics.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data obtained from the study is confidential.
Conflicts of Interest: The authors declare no conflict of interest. The funders agreed on the publication
of the results.
Appl. Sci. 2022, 12, 3356 9 of 9
Abbreviations
The following abbreviations are used in this manuscript:
References
1. Eldred, K.M. Acoustic Loads Generated by the Propulsion System; Technical Report; National Aeronautics and Space Administration:
Washington, DC, USA, 1971.
2. Jiang, C.; Han, T.; Gao, Z.; Lee, C.H. A review of impinging jets during rocket launching. Prog. Aerosp. Sci. 2019, 109, 100547.
[CrossRef]
3. Arenas, J.P.; Margasahayam, R.N. Noise and vibration of spacecraft structures. Ingeniare. Rev. Chil. Ing. 2006, 14, 251–264.
[CrossRef]
4. Nonomura, T.; Nakano, H.; Ozawa, Y.; Terakado, D.; Yamamoto, M.; Fujii, K.; Oyama, A. Large eddy simulation of acoustic
waves generated from a hot supersonic jet. Shock Waves 2019, 29, 1133–1154. [CrossRef]
5. Tam, C.K. Supersonic jet noise. Annu. Rev. Fluid Mech. 1995, 27, 17–43. [CrossRef]
6. Hoyas, S.; Oberlack, M.; Kraheberger, S.; Álcantara-Ávila, F.; Laux, J. Wall turbulence at high friction Reynolds numbers. Phys.
Rev. Fluids 2021, to appear.
7. Gusman, M.; Housman, J.; Kiris, C. Best Practices for CFD simulations of launch vehicle ascent with plumes-overflow perspective.
In Proceedings of the 49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition,
Orlando, FL, USA, 4–7 January 2011; p. 1054.
8. McGuirk, J. Propulsive jet aerodynamics and aeroacoustics. Aeronaut. J. 2022, 126, 2–58. [CrossRef]
9. Tatsukawa, T.; Nonomura, T.; Oyama, A.; Fujii, K. Multi-objective aeroacoustic design exploration of launch-pad flame deflector
using large-eddy simulation. J. Spacecr. Rocket. 2016, 53, 751–758. [CrossRef]
10. Xing, C.; Le, G.; Shen, L.; Zhao, C.; Zheng, H. Numerical investigations on acoustic environment of multi-nozzle launch vehicle
at lift-off. Aerosp. Sci. Technol. 2020, 106, 106140. [CrossRef]
11. Chin, C.; Li, M.; Harkin, C.; Rochwerger, T.; Chan, L.; Ooi, A.; Risborg, A.; Soria, J. Investigation of the Flow Structures in
Supersonic Free and Impinging Jet Flows. J. Fluids Eng. 2013, 135, 031202. [CrossRef]
12. Camussi, R.; Di Marco, A.; Stoica, C.; Bernardini, M.; Stella, F.; De Gregorio, F.; Paglia, F.; Romano, L.; Barbagallo, D. Wind tunnel
measurements of the surface pressure fluctuations on the new VEGA-C space launcher. Aerosp. Sci. Technol. 2020, 99, 105772.
[CrossRef]
13. Escarti-Guillem, M.S.; Hoyas, S.; García-Raffi, L.M. Rocket plume URANS simulation using OpenFOAM. Results Eng. 2019,
4, 100056. [CrossRef]
14. Wilcox, D.C. Turbulence Modeling for CFD; DCW Industries: La Canada, CA, USA, 1998; Volume 2.
15. Menter, F. Zonal two equation kw turbulence models for aerodynamic flows. In Proceedings of the 23rd Fluid Dynamics,
Plasmadynamics, and Lasers Conference, Orlando, FL, USA, 6–9 July 1993; p. 2906.
16. Ramírez, F.N.; Escarti-Guillem, M.S.; García-Raffi, L.M.; Hoyas, S. A study of the mesh effect on a rocket plume simulation.
Results Eng. 2022, 13, 100366. [CrossRef]
17. Kraposhin, M.V.; Banholzer, M.; Pfitzner, M.; Marchevsky, I.K. A hybrid pressure-based solver for nonideal single-phase fluid
flows at all speeds. Int. J. Numer. Methods Fluids 2018, 88, 79–99. [CrossRef]
18. Mortain, F.; Cléro, F.; Palmieri, D. Full Scale Acoustic Source Identification on VEGA Launch Pad at Lift-Off. In Proceedings of
the ICSV26, Montreal, QC, Canada, 7–11 July 2019.
19. Secretariat, E. Spacecraft mechanical loads analysis handbook. In European Cooperation for Space Standardization; ESA Requirements
and Standards Division: Noordwijk, The Netherlands, 2013.