Sie sind auf Seite 1von 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/6313909

Introduction. Computational aerodynamics

Article  in  Philosophical Transactions of The Royal Society A Mathematical Physical and Engineering Sciences · November 2007
DOI: 10.1098/rsta.2007.2014 · Source: PubMed

CITATIONS READS
3 1,259

1 author:

P. G. Tucker
University of Cambridge
380 PUBLICATIONS   3,066 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Jet Engine Intakes View project

LES for Turbines View project

All content following this page was uploaded by P. G. Tucker on 10 July 2014.

The user has requested enhancement of the downloaded file.


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

Introduction. Computational aerodynamics


Paul G Tucker

Phil. Trans. R. Soc. A 2007 365, doi: 10.1098/rsta.2007.2014, published 15


October 2007

References This article cites 19 articles, 1 of which can be accessed free


http://rsta.royalsocietypublishing.org/content/365/1859/2379.ful
l.html#ref-list-1

Email alerting service Receive free email alerts when new articles cite this article - sign up
in the box at the top right-hand corner of the article or click here

To subscribe to Phil. Trans. R. Soc. A go to:


http://rsta.royalsocietypublishing.org/subscriptions
Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

Phil. Trans. R. Soc. A (2007) 365, 2379–2388


doi:10.1098/rsta.2007.2014
Published online 22 May 2007

Introduction. Computational aerodynamics


B Y P AUL G. T UCKER *
Civil and Computational Engineering Centre, University of Wales, Swansea,
Singleton Park, Swansea SA2 0QG, UK

The wide range of uses of computational fluid dynamics (CFD) for aircraft design is
discussed along with its role in dealing with the environmental impact of flight. Enabling
technologies, such as grid generation and turbulence models, are also considered along
with flow/turbulence control. The large eddy simulation, Reynolds-averaged Navier–
Stokes and hybrid turbulence modelling approaches are contrasted. The CFD prediction
of numerous jet configurations occurring in aerospace are discussed along with
aeroelasticity for aeroengine and external aerodynamics, design optimization, unsteady
flow modelling and aeroengine internal and external flows. It is concluded that there is a
lack of detailed measurements (for both canonical and complex geometry flows) to
provide validation and even, in some cases, basic understanding of flow physics. Not
surprisingly, turbulence modelling is still the weak link along with, as ever, a pressing
need for improved (in terms of robustness, speed and accuracy) solver technology, grid
generation and geometry handling. Hence, CFD, as a truly predictive and creative design
tool, seems a long way off. Meanwhile, extreme practitioner expertise is still required and
the triad of computation, measurement and analytic solution must be judiciously used.
Keywords: turbulence; flow control; CFD; aeroelasticity; aeroengines; unsteady flow

1. Introduction

Nowadays, the uses of computational fluid dynamics (CFD) in aerospace


engineering are vast, extending beyond traditional design conditions (Spalart &
Bogue 2003) to numerous peripheral components and aerospace aspects. These
include cabin ventilation and noise (also encompassing outflow valves and their
acoustic influence), avionics/electronics, combustion and fire management. Also,
there are environmental issues such as wake vortices, during and prior to take-off
the resilience of craft/engines to airport environment (Strelets 2001) and engine
ground vortices (Yadlin & Shmilovich 2006). The latter self-generated vortices
can cause catastrophic engine damage (this is mainly owing to large debris
ingestion). Owing to space constraints, it is not possible to cover all of these
aspects here. Furthermore, the current issue is restricted to fixed wing aircrafts
and excludes rotorcraft or hybrids (Spalart et al. 2003).
The environmental impact of aviation is currently very much to the fore and
the reduction of noise and carbon dioxide emissions is currently a key European
and US aerospace priority. To reign in the latter, the European Parliament is
* p.g.tucker@swansea.ac.uk
One contribution of 9 to a Theme Issue ‘Computational fluid dynamics in aerospace engineering’.

2379 This journal is q 2007 The Royal Society


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

2380 P. G. Tucker

seeking to introduce extra charges (via a quota system similar to the one that
industry is subjected to) starting from 2011 based on carbon dioxide emission
levels. For engines, the noise sources can be broken down into turbomachinery
noise (relating to rotating and stationary blades in close proximity for fans,
turbines and compressors), core noise (due to the combustion process) and jet
noise. For airframe noise, sources are high-lift devices such as slats and flaps and
also the landing gear. Discussion of computational aeroacoustics modelling of the
latter can be found in Hedges et al. (2002) and Souliez et al. (2002). However, this
current issue focuses on jet noise.
A massive proportion of an aircraft’s cost is in the avionics/electronics. With
fly-by-wire, unmanned vehicles, automatic landing and the large amounts of
electronics monitoring contained inside modern gas turbine aeroengines the
application of CFD to electronics/avionics is an area of emerging importance.
The environment is generally hostile. However, CFD application expertise in
electronics/avionics (Tucker 1997) seems to lag that in other aerospace areas and so
this is not addressed here. However, it is sufficient to say that many electronic/
avionic system flows involve non-aerodynamic geometries and consequently exhibit
massive separation and hence serious turbulence modelling challenges (Tucker &
Liu 2005). This is also likely to be true of cabin ventilation flows.
This theme issue has papers in the following areas: turbulence (Gatski et al.
2007); flow/turbulence control (Carpenter et al. 2007); propulsive jet dynamics
and noise (Birch et al.); aeroengine external air systems (Dawes 2007); aeroengine
internal air systems (Chew & Hills 2007); aeroelasticity for aeroengines and
external aerodynamics (Bartels & Sayma 2007); design optimization (Keane &
Scanlan 2007); and unsteady flow modelling (Hassan et al. 2007).

2. Turbulence

Obviously, at the high Reynolds numbers experienced with most aerospace flows,
turbulence modelling is a key issue and so the review of Gatski et al. (2007) is
dedicated to this important area. The modelling of complex time-varying turbulent
structures that involve a large range of temporal and spatial scales (the smallest of
which is the Kolmogorov microscale) has always been the weak link in the numerical
prediction of fluid flows. This is perhaps not surprising since (although the Navier–
Stokes do provide a complete description of turbulence) as Richard Feynman said
‘Turbulence is the last great unsolved problem in classical physics’.
The move to explore off-design conditions such as aircraft at take-off and landing,
where there are flaps at high angles of deflection and hence massively separated
flows, has recently drawn greater attention to turbulence models and their
inadequacies (Spalart & Bogue 2003; Tucker 2006). The large-scale, geometry-
dependent eddies arising from massively separated flows ideally need special eddy-
resolving turbulence modelling treatments. However, to resolve scales approaching
the Kolmogorov inertial sub-range requires especially fine numerical grids and
ideally higher-order schemes applicable to unstructured meshes. Eddy-resolving
strategies that reduce grid demands (through incorporating a degree of modelling)
such as large eddy simulation (LES) and detached eddy simulation (DES; Spalart
1999) are mostly discussed in the article by Secundov et al. (2007). With LES, the
larger eddies are resolved and the smaller ones modelled. DES is similar but full

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

Introduction 2381

Reynolds-averaged Navier–Stokes (RANS) modelling is used near walls, i.e. it is a


hybrid LES–RANS method. Eddy-resolving methods are potentially important for
the exploration of approaches to the reduction of turbulence-generated noise where
the delicate control of turbulence interactions is necessary as well as the accurate
prediction of turbulence information. However, the key questions are as follows:
when will computer power be sufficient for their practical design use and meanwhile
what is their value? Secundov et al. (2007) grapple with this question.
The review of Gatski et al. (2007) gives the state-of-the-art and RANS
development priorities along with much background information on RANS
modelling. This turbulence modelling approach is still a mainstay of aerospace
design and will, according to estimates, be so for many years to come. Indeed, it
seems important to note that the design of an aircraft requires many thousands of
simulations and so in many aerospace companies solution of even the RANS
equations is relatively rare with Euler or even full potential flow solutions (Samant
et al. 1987) being made. Even the latter at design conditions can predict
incremental drag for geometry changes with acceptable accuracy (Tinoco 2001).
However, it should be stressed that incremental changes are considerably easier to
quantify than actual drag values. However, knowledge of the latter is vital if the
designed aeroplane is to meet its design specifications and not end up languishing
in a desert scrap heap, hence RANS will soon dominate simpler design methods.
The averaging processes, necessary to give tractable solutions, in both RANS and
LES are compared by Gatski et al. (2007); the discussion includes the use of
temporal averaging in LES (standard LES uses averaging in space). Such
information is important for eventually conceptualizing sound hybrid LES–RANS
methods. Currently, these are showing promising results (albeit at considerable
computational expense) for flows where RANS models fail. However, the
interfacing of time- and space-averaged equations, although easy, is theoretical
questionable. Hence, the detailed discussion is not just helpful in setting out the
foundations for RANS models. Gatski identifies the following current challenges:
transition modelling, relaminarization, separation and flow control. It perhaps
seems doubtful with the complex physics involved in these processes that RANS
modelling will ever be adequate. As noted by Spalart & Bogue (2003), unsteady
RANS (as with virtually all other applications) is unlikely to be of use for flow
control. The fascinating issue of multiple solutions arising as a result of the chosen
RANS model is also briefly addressed by Gatski et al. (2007). Further discussion
on such issues can be found in Rumsey (2006) and Rumsey et al. (2006).

3. Flow/turbulence control

Turbulence governs airframe drag and so to reduce environmental emissions


(from excessive fuel burn) clearly, it is important to control it. Carpenter et al.
(2007) outline the various drag sources and their relative importance.
Aerodynamics design has improved so much that now over 50% of aircraft
drag is due to skin friction, hence there is a pressing need to reduce it. Intricate,
elongated near-wall structures (sublayer streaks) play a key role in the
generation of skin friction. Hence, a novel numerical model for sublayer streak
prediction is presented by Carpenter et al. (2007). Also, practical aspects relating
to the control of near-wall structures are discussed along with elements of flow

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

2382 P. G. Tucker

control in general. The use of microelectromechanical systems (MEMS) and


other active control devices in aerodynamic flow control (for both the propulsion
units—especially with respect to jet noise—and airframes) is finding increasing
interest. Hence, the assessment of the future potential of such technologies
through numerical modelling is important. A deeper understanding of near-wall
behaviour and streak generation is also likely to be of importance in improving
the theoretical foundations of RANS, hybrid LES–RANS and pure LES methods,
thus adding to the importance of the work of Carpenter et al. (2007). As noted by
Pope (2004), at cruise, a Boeing 777 wing has approximately 108 streaks (these
estimates probably do not include the effects of pressure gradient and sweep)
presenting a key LES resolution issue. Similar stark statistics relating to spatial
and temporal scales, but this time for the fuselage of an Airbus A340-300, are
brought out by Carpenter et al. (2007). According to their estimates for the
Airbus 340-300, there are 2000 streaks mK1 and approximately 20!108 for the
whole fuselage. Indeed, it is primarily owing to these structures that Spalart &
Bogue (2003) estimate it will not be possible until 2050 to carry out a LES of a
full airframe and direct solution of the Navier–Stokes equations without recourse
to turbulence modelling will have to wait until 2080. It is these daunting
statistics that the hybrid LES–RANS methods are intended to overcome. This
being achieved by covering the streaks with a RANS model.
The projected demand for air transport suggests a doubling of the world aircraft
fleet by 2020. Hence, it is becoming urgent to take steps to reduce the environmental
impact of take-off noise. In the worst case, this can be more than just annoying. It is
potentially a contributory factor towards illnesses such as hypertension (Rosenlund
et al. 2001). Also, continual reductions in permitted take-off noise levels are placing
the commercial viability of more established aircraft designs in jeopardy. At take-
off, a major noise source is the propulsive jet. The article by Secundov et al. (2007)
largely focuses on the RANS modelling of various jets and the prediction of far-field
noise along with the escape trajectories of acoustic rays from within jets.

4. Propulsive jet dynamics and noise

Secundov et al. (2007) outline the daunting task of using RANS to model the
substantially different flow physics for isothermal, non-isothermal, sub- and
supersonic jets for the range of complex nozzle geometries found in different
aerospace contexts. Modelling challenges identified are twin nozzle rocket
plumes, jets with co-flow and wall jets. Study of the last relates to jet blast, i.e.
the damage to cargo, small craft and buildings from propulsive jets. Co-flowing
jets occur in commercial aircraft engines. The surprising lack of understanding of
the flow physics of even basic jets is outlined.
Most importantly, Secundov and colleagues show that for jets with co-flow, the
flow structure is very sensitive (with respect to concentricity) to inflow conditions
potentially rendering much of the past data for this configuration invalid. It is well
known that jets are highly sensitive to external environmental conditions but the
concentricity aspect is new. The addition of a pylon, which is representative of the
real engine operating geometry, improves the situation. Hence, this aligns with
the work of Dawes (2007) in this issue, which strongly advocates the need to model
real geometry influences. Spalart & Bogue (2003) also raise this matter.

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

Introduction 2383

Secundov and colleagues highlight RANS models, which can be of value; for
example, in scenarios where the global flow structure complexity is pressure and
not turbulence stress driven. Also, the importance of appropriately modelling
inflow conditions is noted and the general incomplete information on this with
regard to measurements is highlighted. Hence, it appears that more work on
creating adequate validation data is required, with perhaps both experimental-
ists and CFD practitioners being brought together at the outset of new
experimental programs. In this way, it could be ensured that the measured data
have adequately defined boundary conditions for the meaningful validation and
refinement of numerical models. Indeed, preliminary CFD simulations could
obviously contribute to the design of the experiments.
Secundov and colleagues also discuss the use of LES in a jet noise context. LES
results incorporate a new use of the modified Eikonal equation to explore how
instantaneous velocity (which vary dramatically from the mean) and temperature
gradients deflect acoustic rays. The Eikonal equation is an approximation to the
wave equation for short waves. Acoustic waves are shown to be dramatically
deflected (in relation to mean velocity and density gradients) by turbulence. The
turbulence is shown to make the jet opaque to short waves. Hence, acoustic
sources in one quadrant do not produce acoustic rays in others.
With regard to industrial use of LES, it is concluded that issues of problem
definition (especially with respect to inflow boundary conditions), solution
uniqueness, near-wall modelling, transition issues and grid structure along with its
implications on filtering (to achieve spatial averaging) are of potentially greater
importance for practical jet flows than the LES model itself. The latter being
theoretically questionable. This conclusion implies that much time is spent on the
implicit LES (where the numerical discretization partly acts as an LES model; see
Boris et al. 1992) versus LES debate. It seems focused that LES simulations should
provide flow physics insights and help with RANS model development. The value of
LES at impacting on the design context more directly as yet remains unclear.
The full coupling of the structural dynamics, pilot input/response (Spalart &
Bogue 2003) and also manufacturing constraints with aerodynamic design is an
ultimate future goal. This coupling allows quicker, safer and more innovative
aerospace designs.

5. Aeroelasticity for aeroengines and external aerodynamics

Bartels & Sayma (2007) present an extensive review on the state-of-the-art for both
aeroelasticity in airframe and turbomachinery design. NASA’s recent mandate to
deliver payloads to Mars and ultimately return samples to Earth has brought about
the need to model highly flexible, inflatable thin membrane decelerators,
reminiscent of parachutes. This has motivated the need for aeroelasticity studies
that can model highly nonlinear structures with nonlinear material properties. The
high temperature of re-entry needs coupling of the thermal fields in the fluid flow
and structure. Full validation of such simulations presents a significant challenge.
However, Bartels & Sayma (2007) report encouraging agreement for the measured
maximum deflection at a high Mach number. This work perhaps represents
the state-of-the-art in nonlinear aeroelasticity modelling. Most aeroelasticity
validations involve lower-order quantities.

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

2384 P. G. Tucker

A fuller comparison of model performance assessment with actual flow field


(mean velocities and turbulence statistics) and surface heat transfer data would
help to better diagnose model deficiencies.
Presumably, inflatable decelerators exhibit massively separated flow. Hence,
ideally these aeroelasticity studies should make use of hybrid LES–RANS
techniques. However, due to computational expense, aeroelasticity work using
this turbulence modelling approach seems to be rare. Bartels & Sayma (2007)
discuss the use of reduced order models as well as the extreme of full nonlinear
aeroelasticity models. These will form an essential ingredient in more routine design
studies, computer-driven design optimization procedures and perhaps models of the
way pilots interact with aircraft. With design optimization, for given targets and
constraints (including, for example, those arising from the manufacturing process
and also component weights), mathematical models coupled with CFD are used to
seek optimal solutions. These are now an important part of aerospace engineering.

6. Design optimization

The aerospace design-led optimization paper by Keane & Scanlan (2007) gives
some idea of the daunting task the aerospace design engineer faces. Many
challenges arise from the numerous conflicting design requirements. Keane &
Scanlan (2007) review design search and optimization approaches. A multi-
objective, multidisciplinary example for a transonic civil aircraft design is given.
The analysis encompasses structural integrity, weight, manufacturing cost and
aerodynamic performance. Strength and weight are modelled using semi-
empirical schemes based on aerospace company practice. Costing is component
based and considers elements such as spars, ribs and skins. The analysis requires
the automated linking of numerous different codes. Hence, presumably good
scripting skills are needed. Despite the numerous elements (CAD engine, mesh
generator, analysis codes, post-processing and optimization tools) and the CFD
solution being Euler based, it still proved to be the most expensive element
needing about half the total solution time. Perhaps not unexpectedly, around
50% of CFD solutions in the design optimization cycle failed. This shows the
need for improved automatic mesh generation and CFD solver technology.
Clearly, the use of advanced turbulence modelling strategies (except for more
basic design optimizations) is some way off—the designer, needing to balance so
many things, is always forced to use methods that are less expensive, and thus
less accurate, than the pure analyst. It seems probable that computational-based
optimization tools, in an environment with many diverse requirements that must
be held in tension, offer a rational basis for design decision making, which must
make a very positive impact on the design process.

7. Unsteady flow modelling

Hassan et al. (2007) discuss techniques for solving unsteady flow equations with
moving unstructured meshes used to resolve either flow features or boundary
movement. As might be expected, such problems present the major challenge of
maintaining a valid mesh. Approaches suitable for problems with extreme
boundary movement are discussed. To remove the serial bottlenecks discussed by

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

Introduction 2385

Dawes (2007) (see §8), these involve parallel local remeshing. Such approaches
are important for store release, formation flying (Aftosmis et al. 1997), air
refuelling, extreme control surface movement, aeroelasticity for highly flexible
structures (as discussed by Bartels & Sayma (2007)) and design optimization—
when the development of radically different design solutions is sought. Flow-
based adaptation potentially lends itself to LES and hence appears to be an
important emerging area. Explicit and implicit temporal integration schemes are
contrasted. New approaches for alleviating the severe Courant–Friedrichs–Lewy
condition constraint for explicit schemes are discussed. Again, the question of
whether the potentially more efficient, for certain classes of problems, explicit
schemes can be implemented is strongly linked to adequate mesh control.
However, clearly an implicit solver is most desirable. Also, spatial truncation
errors will interact with those from the temporal schemes. Hence, these two
elements should be carefully selected to ensure that they work in harmony and
are compatible with the class of problem being considered.

8. Aeroengine external air systems

Dawes (2007) charts the historical evolution of CFD analysis for external
systems related to aeroengine flows (i.e. blade-related flows). This spans simple
idealized geometries to the current status where, to affect engine performance,
real geometry (elements of geometry that have largely been ignored in the past)
influences must be taken into account. According to Spalart & Bogue (2003), real
geometry modelling is also essential for airframe aerodynamics. Dawes (2007)
presents a grand vision utilizing technology from the computer games industry.
The work gets to grips with the key issues involved in taking CFD from being an
analysis tool to a true aerodynamics design tool that does not stifle creative
activity. The use of CFD in the games/graphics industries is more advanced than
one might expect and the industry now provides much advanced computational
modelling technology. With the initial use of body-fitted grids in turbomachinery
engineering, boundary representation geometry models were perfectly adequate.
However, with the drive towards modelling the full geometry, such approaches
make geometry modification slow with CAD geometry clean-up (to get
watertight geometry) taking a substantial time. Hence, the case for moving
from explicit to implicit geometry is made, the use of signed distance fields for
the latter being advocated. Indeed, this distance field is also necessary in many
robust turbulence models and can, as noted by Dawes (2007), be used as a basis
for generating high-quality surface meshes. As noted by Spalart & Bogue (2003),
with a mind on turbulence modelling, the capability to model real geometry will
enable a clearer picture of the capabilities of turbulence models for certain
complex geometry applications. Parallel processing (and partly memory-based)
bottlenecks, arising from inherently serial elements such as geometry modelling,
mesh generation and partitioning, post-processing and data exchange between
different packages, are discussed by Dawes (2007). With reasonably effective use
now being made of parallel processing for solution of the flow equations, the need
to remove these bottlenecks is becoming more apparent. This point, along with
the need to model real geometry and hence use unstructured solvers, is echoed by
Chew & Hills (2007).

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

2386 P. G. Tucker

9. Aeroengine internal air systems

Chew & Hills (2007) focus on aeroengine internal air systems. Since up to 5% of the
engine’s power can go into supporting these systems, through, for example,
supplying blade cooling, they are of significant importance. These enclosed rotating
system flows exhibit complex physics with many aspects being reminiscent of
geophysical flows. LES has its origins in such flows. It is clear from this paper that,
for internal air systems, LES holds promise. Generally, due to the complex flow
physics, RANS models perform poorly for these flows. The need for robust meshing
(and automatic adaptation) especially in a design optimization context is again
identified. The internal air systems can have significant geometrical complexity. It
is clear that design optimization is beginning to play a role in internal air systems
and helping with the practical uptake of design improvements. The need for greater
automation of the CFD process to reduce user dependence on results is identified.
In the multiphysics, internal air systems environment, where the thermal coupling
of the solid and fluid zones is important, robust automation becomes even more
important. The internal air systems discussed by Chew & Hills (2007) feed passages
inside turbine blades that provide blade cooling. The computation of these flows,
which again present severe turbulence and geometry modelling challenges, is
discussed by Iacovides & Raise (1999).

10. CFD and measurements

For much of the work discussed above, it appears that there is a lack of detailed
measurements to provide validation and even a basic understanding of the flow
physics. This makes serious assessment of turbulence models and, in some cases,
development impossible. Frequently, for aerodynamics configurations, wind-
tunnel data just consist of global data on lift and drag coefficients with perhaps
surface pressure measurements. Also, wind-tunnel influences can be excessive.
This is especially a problem when studying high-lift configurations where severe
flap deflections can generate large vortices adjacent to the wind-tunnel wall.
However, for serious assessment and refinement of turbulence model per-
formance, turbulence statistics data, free from wind-tunnel wall interference, are
necessary. For jets, inflow conditions are generally inadequately defined, the
measurements being focused on the actual jet flow away from the outlet. Again,
this makes serious assessment of numerical modelling extremely difficult. Also,
for jets with co-flow, it seems to have taken a surprisingly long time for the
sensitivity of the flow-to-inflow concentricity to come to light. This must be due
to experimentalists restricting measurements to just one plane and erroneously
taking axisymmetric flow for granted. This is also, in a sense, reminiscent of
internal air systems in gas turbine aeroengines. For many years, axisymmetric
flow was assumed and complex flow instability mechanisms ignored. Impor-
tantly, when wind-tunnel tests are repeated, shock locations can vary by 20% of
the wing chord. However, repeat measurements are seldom made. Hence, these
‘uncertainty data’ cannot be supplied to the CFD practitioners. In the light of
this, it seems surprising, considering the potential data scatter, that most CFD
simulations seem to capture shock location very well. The surprising level of
agreement perhaps reflects the fact that CFD is still a heavily postdictive

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

Introduction 2387

process. With many applications, the CFD practitioner has to make a wide range
of modelling assumptions. Through these, solution control can be exerted. This
appears especially true for the prediction of jet noise.
At perhaps a more fundamental level, it has been fairly widely assumed that the
Kármán constant (k), used in Prandtl’s mixing-length expression, is universal,
having an established value of approximately 0.41. However, recent evidence
suggests that the ‘constant’ is more dependent on the nature of the boundary layer
with values in the range 0.38!k!0.45 (George 2006). Spalart (2006) notes that for
modern passenger aircraft, a 2% k decrease will cause a 1% decrease in predicted
drag. This difference is of extreme importance. If the aircraft, which is initially sold
and manufactured based on CFD, does not achieve the expected range, then the
consequences for the manufacturer could be disastrous. Hence, even basic boundary
layer measurements to establish key CFD code constants are of high importance. As
noted by George (2006), improved wind-tunnel facilities are necessary. Then
boundary layer measurements for flows that have sufficiently high Reynolds
numbers to be considered truly turbulent can be made. The same is true for jets for
which, when heated, the Reynolds number can drop fast. Much of the available jet
flow and noise data, particularly at lower Mach numbers, has been taken at relatively
low Reynolds numbers, and so this data seems suspect. This is important—the
effective Mach number (based on the velocity difference between the primary and fan
streams) of hot jets encountered in real airplane applications can be low. Hence, it
seems important to create high-quality datasets, both for canonical flows and
complex geometries. Currently, the trend unfortunately seems to be for the data
quality to decrease with increasing realism in geometrical complexity and testing.
I had encouraged authors to be prophetic in this theme issue but sadly they
(perhaps wisely) refrained. Chapman et al. (1975) predicted the demise of the wind
tunnel in around 1985. However, turbulence is still the weak link. Both understanding
and directly predicting turbulence needs powerful computers and the ability to
visually interpret the data produced. Processor power increases are being strongly
driven by the children’s computer games industries and even computer graphics
(technology in this area being driven by blockbuster movies such as Jurassic Park).
Perhaps the next generation will have access to truly predictive CFD technology.
Meanwhile, extreme practitioner expertise is still required and the triad of
computation, measurement and analytic solution must be judiciously used.

References
Aftosmis, M. J., Berger M. J. & Melton J. E. 1997 Robust and efficient mesh generation for
component based geometry. AIAA Paper 97-0196, 1–13.
Bartels, R. E. & Sayma, A. I. 2007 Computational aeroelastic modelling of airframes and
turbomachinery: progress and challenges. Phil. Trans. R. Soc. A 365, 2469–2499. (doi:10.1098/
rsta.2007.2018)
Boris, J. P., Grinstein, F. F., Oran, E. S. & Kolbe, R. L. 1992 New insights into large eddy
simulation. Fluid Dyn. Res. 10, 199–228. (doi:10.1016/0169-5983(92)90023-P)
Carpenter, P. W., Kudar, K. L., Ali, R., Sen, P. K. & Davies, C. 2007 A deterministic model for the
sublayer streaks in turbulent boundary layers for application to flow control. Phil. Trans. R.
Soc. A 365, 2419–2441. (doi:10.1098/rsta.2007.2016)
Chapman, D. R., Mark, H. & Pirtle, M. W. 1975 Computers vs wind tunnels for aerodynamic flow
simulations. Astronaut. Aeronaut. 13, 22–35.

Phil. Trans. R. Soc. A (2007)


Downloaded from rsta.royalsocietypublishing.org on March 12, 2014

2388 P. G. Tucker

Chew, J. W. & Hills, N. J. 2007 Computational fluid dynamics for turbomachinery internal air
systems. Phil. Trans. R. Soc. A 365, 2587–2611. (doi:10.1098/rsta.2007.2022)
Dawes, W. N. 2007 Turbomachinery computational fluid dynamics: asymptotes and paradigm
shifts. Phil. Trans. R. Soc. A 365, 2553–2585. (doi:10.1098/rsta.2007.2021)
Gatski, T. B., Rumsey, C. L. & Manceau, R. 2007 Current trends in modelling research for turbulent
aerodynamic flows. Phil. Trans. R. Soc. A 365, 2389–2418. (doi:10.1098/rsta.2007.2015)
George, W. K. 2006 Emerging experiments for CFD session. In 36th Fluid Dynamics Conf. and
Exhibit, 5–8 June, San Francisco, CA.
Hassan, O., Morgan, K. & Weatherill, N. 2007 Unstructured mesh methods for the solution of the
unsteady compressible flow equations with moving boundary components. Phil. Trans. R. Soc. A
365, 2531–2552. (doi:10.1098/rsta.2007.2020)
Hedges, L. S., Travin, A. K. & Spalart, P. R. 2002 Detached-eddy simulations over a simplified
landing gear. J. Fluids Eng. Trans. ASME 124, 413–423. (doi:10.1115/1.1471532)
Iacovides, H. & Raise, M. 1999 Recent progress in the computation of flow and heat transfer in
internal cooling passages of turbine blades. Int. J. Heat Fluid Flow 20, 320–328. (doi:10.1016/
S0142-727X(99)00018-1)
Keane, A. J. & Scanlan, J. P. 2007 Design search and optimization in aerospace engineering.
Phil. Trans. R. Soc. A 365, 2501–2529. (doi:10.1098/rsta.2007.2019)
Pope, S. B. 2004 Ten questions concerning the large-eddy simulation of turbulent flows. New J.
Phys. 6, 1–24. (doi:10.1088/1367-2630/6/1/035)
Rosenlund, M. et al. 2001 Increased prevalence of hypertension in a population exposed to aircraft
noise. Occup. Environ. Med. 58, 769–773. (doi:10.1136/oem.58.12.769)
Rumsey, C. L. 2006 Apparent transition behaviour of widely-used turbulence models. In 36th AIAA
Fluid Dynamics Conf. and Exhibit, 5–8 June, San Francisco, California, AIAA 2006-3906.
Rumsey, C. L., Pettersson Reif, B. A. & Gatski, T. B. 2006 Arbitrary steady-state solutions with
the K-epsilon model. AIAA J. 44, 1586–1592.
Samant, S. S. et al. 1987 TRANAIR: a computer code for transonic analyses of arbitrary
configurations. AIAA Paper 87-0034, pp. 1–18.
Secundov, A. N., Birch, S. F. & Tucker, P. G. 2007 Propulsive jets and their acoustics. Phil. Trans.
R. Soc. A 365, 2443–2467. (doi:10.1098/rsta.2007.2017)
Souliez F. J., Long L. N., Morris P. J. & Sharma A. 2002. Landing gear aerodynamics noise
prediction using unstructured grids. In 40th AIAA Aerospace Sciences Meeting and Exhibit,
14–17th January, Reno, NV, AIAA 2002-0799.
Spalart, P. R. 1999 Strategies for turbulence modelling and simulations. In Proc. 4th Int. Symp. on
Engineering Turbulence Modelling and Measurements, 24–26 May Ajaccio, Corsica, France.
Spalart, P. 2006 Turbulence. Are we getting smarter? Fluid Dynamics Award Lecture. In 36th
Fluid Dynamics Conf. and Exhibit, 5–8 June, San Francisco, CA.
Spalart, P. R. & Bogue, D. R. 2003 The role of CFD in aerodynamics, off-design. Aeronaut. J. 107,
323–329.
Spalart, P., Hedges, L., Shur, M. & Travin, A. 2003 Simulation of active flow control on a stalled
airfoil. Flow Turbul. Combust. 71, 361–373. (doi:10.1023/B:APPL.0000014925.91304.42)
Strelets, M. 2001 Detached eddy simulation of massively separated flows. In 39th AIAA Aerospace
Sciences Meeting and Exhibit, 8–11 January, Reno, NV, AIAA 2001-0879.
Tinoco, E. N. 2001 An assessment of CFD prediction of drag and other longitudinal characteristics.
In 39th Aerospace Sciences Meeting and Exhibit, Reno, NV, 8–11 January. AIAA-2001-1002.
Tucker, P. G. 2006 Turbulence modelling for problem aerospace flows. Int. J. Numer. Methods
Fluids 51, 261–283. (doi:10.1002/fld.1120)
Tucker, P. G. 1997 CFD applied to electronic systems: a review. Trans. IEEE (CPMTA) 20, 518–529.
Tucker, P. G. & Liu, Y. 2005 Contrasting CFD for electronic systems modeling with that for
aerospace. In Proc. 6th IEEE EuroSimE Conf., Berlin, Germany, 18–20th April (eds L. J. Ernst
et al.), pp. 618–625.

Phil. Trans. R. Soc. A (2007)

View publication stats

Das könnte Ihnen auch gefallen