Sei sulla pagina 1di 24

FL44CH21-Alonso ARI 21 November 2011 13:0

ANNUAL
REVIEWS Further
Multidisciplinary Optimization
Click here for quick links to
Annual Reviews content online, with Applications to
including:
Other articles in this volume
Top cited articles
Sonic-Boom Minimization
Top downloaded articles
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

Our comprehensive search


Juan J. Alonso and Michael R. Colonno
Department of Aeronautics and Astronautics, Stanford University, Stanford, California 94305;
email: jjalonso@stanford.edu, mcolonno@stanford.edu
by Brown University on 12/14/12. For personal use only.

Annu. Rev. Fluid Mech. 2012. 44:50526 Keywords


First published online as a Review in Advance on supersonic ow, perceived loudness, boom propagation, robust design,
November 2, 2011
noise, aircraft design, supersonic design
The Annual Review of Fluid Mechanics is online at
uid.annualreviews.org Abstract
This articles doi: This article presents a review of key historical contributions, the current sta-
10.1146/annurev-uid-120710-101133
tus, and future research avenues in support of the development of supersonic
Copyright  c 2012 by Annual Reviews. aircraft that are sufciently quiet so that they can be allowed to y superson-
All rights reserved
ically over land. For this goal to be achievable, in addition to overcoming
0066-4189/12/0115-0505$20.00 many other challenges in aerodynamics, structures, propulsion, acoustics,
and aeroservoelasticity, the pressure signature created by the aircraft must
be such that, when it reaches the ground, (a) it can barely be perceived by
the human ear, and (b) it results in disturbances to man-made structures that
do not exceed the threshold of annoyance for a signicant percentage of the
population. In other words, the ground-boom signature must meet a number
of key constraints that can be appropriately quantied. In designing aircraft
with low sonic booms, it is important to understand (a) how pressure dis-
turbances are generated and how they propagate through the atmosphere,
(b) under which conditions will the pressure signature created by an aircraft
evolve to generate an acceptable low-boom signature at the ground, and
(c) what multidisciplinary trade-offs need to be made to realize low-boom
aircraft that are also economically and environmentally compliant. This ar-
ticle discusses each of these areas separately, assesses the accomplishments in
each topic, identies signicant shortcomings, and suggests future research
efforts (some already ongoing) that have the potential to yield solutions to
all these issues.

505
FL44CH21-Alonso ARI 21 November 2011 13:0

1. INTRODUCTION AND BACKGROUND


Any aircraft that ies supersonically through the atmosphere creates pressure disturbances that
result both from the displacement of a volume of uid as the aircraft passes by and from the
generation of the aerodynamic lift required to adhere to a desired ight path. These pressure
disturbances propagate in all directions, and a signicant portion of the acoustic energy reaches
the ground despite typical cruise altitudes in the neighborhood of 60,000 ft. At the ground, these
pressure disturbances reect with some level of impedance and are perceived as what we call a
sonic boom. Although it is entirely plausible that sonic booms had been perceived prior to the
advent of supersonic ight (with the crossing of the sound barrier by the experimental Bell X-1
aircraft in 1947), these booms would have resulted from supersonic projectiles with sonic-boom
patterns that were largely unknown. A true awareness of the sonic boom as a physical phenomenon
associated with supersonic ight had not been seriously considered until the early 1950s (Gold
1952, Plotkin 1989, Warren 1953). The understanding of the physical phenomena that lead to
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

ground booms quickly advanced through the work of Whitham (1952), Hayes (1947), McLean
(1965, 1968), Jones (1961), and Seebass, George, and Darden (Darden 1977, Seebass 1969a,b,
Seebass & George 1972). Over a period of approximately 10 years, the physical formation of sonic
booms was understood as the natural nonlinear evolution of the near-eld pressure signature as it
by Brown University on 12/14/12. For personal use only.

propagated away from the aircraft. Some difculties in arriving at the proper physical interpretation
resulted from the fact that, at the time, predictions of the near-eld signature were obtained using
classical linearized supersonic aerodynamics, which does not, to rst order, account for essential
nonlinear phenomena such as waveform steepening and variable-speed wave propagation. Once
these rst-order corrections were incorporated, it was quickly found that the far-eld behavior
of the signatures (which evolved into an N-wave, as can be seen in Figure 1) could be explained
and predicted with relative accuracy. An N-wave refers to the shape of the pressure signature
that results at large distances below the aircraft, as the positive and negative (compressions and

Near field

Mid-field

Far field

Figure 1
Schematic diagram of boom-signal evolution (not to scale). As the near-eld signal propagates away from the
aircraft, various shocks formed by aircraft features coalesce, resulting, in some cases, in the classic N-wave in
the far eld consisting only of a leading and trailing shock.

506 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

expansions with respect to the free stream) pressure disturbances caused by the aircraft slow down
and speed up to make multiple shocks coalesce into leading and trailing shocks separated by a nearly
linear pressure expansion. The name sonic boom results from the fact that, when the aircraft is
small and the leading and trailing shocks of the ground signature are very close to each other, the
human ear perceives a single bang. In reality, and particularly for larger/longer aircraft, the human
ear perceives a sonic boom-boom caused by the leading and trailing shocks with an expansion in
between that has sufciently low-frequency content that is imperceptible to humans.
Considerable activity followed during the 1960s and 1970s as the English and French designed
and developed Concorde, the Russians followed along with the Tupolev TU-144, and the United
States investigated the possibilities of its own supersonic transport. It was quickly realized that,
with the available knowledge at the time, such aircraft would create a signicant sonic boom at
the ground that could not be mitigated. Although one could argue that takeoff and landing-noise
considerations led to the demise of the U.S. supersonic transport and seriously put in jeopardy the
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

limited operations that the Anglo-French Concorde eventually achieved, there is no denying that
sonic-boom considerations led to restrictions on supersonic ight overland, which are enforced
to this day. In fact, on a typical ight from Heathrow to John F. Kennedy Airport, Concorde was
allowed only to reach supersonic speeds once the projected boom carpet fell on the ocean, and
by Brown University on 12/14/12. For personal use only.

it was forced to decelerate to subsonic speeds before arriving at the U.S. East Coast so that no
sonic-boom disturbances would impact populated areas on land. Although Concorde and the TU-
144 were designed without signicant attention to sonic-boom-reduction technologies (Collard
1991), the research community rapidly began to see the possibility of modifying the character of
the sonic-boom signature so that its impact on the ground was minimized.
Sonic-boom minimization became a signicant research activity in the early to mid-1960s,
with McLeans (1965) realization that the asymptotic boom signature (typically an N-wave) takes
several hundred aircraft lengths to be reached in an isothermal atmosphere. The situation improves
further when considering a standard atmosphere, with a nonzero temperature lapse. Previously,
it had been thought that the generation of a far-eld N-wave was inevitable, but with McLeans
realization, it became obvious that one may be able to tailor the near-eld signature so that, by the
time it reaches the ground from a cruise altitude, it is not yet frozen in its far-eld shape, but rather
it is still evolving through what is typically called its mid-eld signature. The mid-eld signature
shape reaching the ground could then be engineered so that its impact could be minimized. For
example (Seebass & Argrow 1998), one might design ground/mid-eld sonic-boom signatures
that, for a given aircraft weight and length, (a) minimize the signature impulse, (b) minimize the
maximum overpressure, and/or (c) minimize the shock pressure rise and use nite rise times for
other portions of the signature. Such approaches to sonic-boom minimization and the kinds of
ground-boom signatures in which they result can be seen in Figure 2.
The main approaches that were proposed to obtain such boom signatures relied on tailor-
ing both the volume and lift distributions of the aircraft (which are responsible for creating the
near-eld signature in the rst place) in such a way that the mid-eld signature exhibited the
characteristics seen in Figure 2. Although other approaches based on Busemann biplane ideas
and energy deposition were proposed in the literature, none has been found to be viable. Because
of the heavy reliance on classical linearized supersonic aerodynamics, the target near-eld pres-
sure signatures were expressed as equivalent area distributions: the area of a body of revolution
(with a blunt base if the lift is nonzero) that, from a signicant distance from the aircraft and at a
particular azimuthal angle, produced the exact same pressure distribution that the original aircraft
would have created (thus the name equivalent). It is fair to say that this seminal work of many
authors resulted in the Jones-Seebass-George-Darden ( JSGD) theory of sonic-boom minimiza-
tion (Darden 1977, Jones 1961, Seebass 1969a,b, Seebass & George 1972), which has provided

www.annualreviews.org MDO for Sonic-Boom Minimization 507


FL44CH21-Alonso ARI 21 November 2011 13:0

a p Is

po

p
b
ps

T
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

Figure 2
(a) Schematic of a ground-boom signature optimized for minimum pressure rise (or overpressure) with
overpressure po and impulse Is . Boom signatures optimized for minimum impulse have a similar form to
by Brown University on 12/14/12. For personal use only.

those optimized for minimum overpressure. (b) Schematic of a ground-boom signature optimized for
minimum leading-shock strength, ps .

target equivalent area distributions for low-sonic-boom designs until presently. Although new
approaches based on nonlinear predictions of the ow around the aircraft and in the near eld are
currently emerging, sonic-boom-minimization approaches still rely on equivalent area distribu-
tions that are similar to those of Seebass. Although Seebass (1969b) was quick to point out that the
impact of sonic booms could also be reduced by leveraging integrated design approaches [which
today we might call multidisciplinary design optimization (MDO)], the majority of his work and
that of others until around 2000 focused on the reduction of the impact of sonic booms alone
through the shaping of the aircraft and its lift distribution. These contributions did not simulta-
neously consider the effects (favorable or detrimental) of/on other disciplines that are integrated
into a viable aircraft design (except in some obvious ways, such as reducing the nose bluntness
that appeared to be required; see Darden 1977). Cheung et al. (1992) pursued some initial holistic
efforts to include other considerations in the design of a low-boom supersonic aircraft within the
context of what was to become the NASA High-Speed Research program. For an aircraft of the
size, range, and cruise Mach number being pursued (300 passengers, Mach 2.4, 5,000 nmi; see
Natl. Res. Council 1997), it became obvious that a low-boom solution did not exist, and the main
aspects of the environmental impact of that aircraft became the impact of high-altitude emissions
on ozone depletion and airport-/community-noise issues (Cheung 1994). However, during the
fall of 2000, DARPA created the Quiet Supersonic Platform program to reassess the available
technologies necessary to develop small supersonic aircraft with sufciently low sonic boom that
they may be allowed to y supersonically over land. The program requirements stated that such an
aircraft would have to be in the 100,000-lb class, y at a cruise Mach number of 2.4 with a range of
6,000 nautical miles, and produce an initial overpressure of less than 0.3 psf. These design re-
quirements were issued as guidelines and were subsequently revised so that the goal would be
achievable. This program (Alonso et al. 2002) represented the rst attempt to build low-boom
requirements into the conceptual design of a complete aircraft that also met all the other necessary
requirements. Efforts by multiple institutions in the Quiet Supersonic Platform program demon-
strated that, for a small aircraft, the requirement of low sonic boom could be made practical. In fact,

508 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

SHAPING GROUND SIGNATURES

The DARPA/NASA/Northrop-Grumman Shaped Sonic Boom Demonstrator (SSBD) project is the only ight-test
activity accomplished to date in which the theory of sonic-boom minimization has been shown to be valid. Because
of the constraints of the underlying F-5E aircraft, the team was able to tailor only the front portion of the signature.
The focus of research now is to show that it is possible to shape the entire signature (including the rear portion at
which multiple shocks arising from the wing, tail surfaces, and propulsion plant interact in fairly complicated ways)
for a truly low-boom design.

Northrop-Grumman, together with DARPA and NASA, went on to modify the front portion of
an F-5E aircraft to demonstrate in ight, for the rst time, that a shaped sonic-boom signature
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

(at least for the front portion of the signature) could be designed into a vehicle and that it could
persist all the way to the ground with varying atmospheric conditions (Pawlowski et al. 2005).
The DARPA/NASA/Northrop-Grumman Shaped Sonic Boom Demonstrator (SSBD) has
allowed us to re-examine the possibilities for aircraft with fully shaped sonic-boom signatures
by Brown University on 12/14/12. For personal use only.

that are signicantly larger in size (75100 passengers at Mach 1.61.8) (see the sidebar, Shaping
Ground Signatures). All the ongoing efforts are carefully looking at the impact of low-boom
requirements on the multidisciplinary design of the entire conguration. Many challenges are
being faced because of the closely integrated nature of the resulting designs and because what is
benecial to reduce the sonic-boom intensity may be detrimental to aerodynamic performance or
the performance of other disciplines (Aronstein & Schueler 2005). Presuming that all technical
challenges can be met, Henne (2005) and Tetzloff & Crossley (2011) have made careful analyses of
the potential markets for both small- and medium-sized low-boom aircraft and have found them
to be of signicant size. Although several efforts are currently ongoing worldwide, the NASA
Supersonics project is currently leading the charge and is pursuing the necessary technologies
required for what it calls N+2 and N+3 low-boom supersonic aircraft (Morgenstern et al. 2010,
Welge et al. 2010), which may one day allow aircraft with signicant potential to impact the
way we travel to materialize. Such aircraft would have both the necessary performance and low
environmental impact that could make them a reality. In addition to efforts pursued by NASA in
the United States, industry participants (including Gulfstream, Boeing, and Lockheed-Martin),
European institutions, and the Japanese JAXA (Chiba et al. 2008) are all currently pursuing ideas for
low-boom aircraft. In some cases, these aircraft are manned, whereas in others they are unmanned
demonstrators ( Jung et al. 2011).
The purpose of this article is not necessarily to review all components of research that can lead
to the realization of integrated, low-boom, supersonic aircraft but rather to highlight ongoing
efforts and suggest research directions that may need to be pursued to make such aircraft a reality.
The following sections provide an overall assessment of each of three areas of research that are
considered to be of fundamental importance, together with references to other review articles or
relevant publications for those who may be interested in examining the theory in more detail.

2. SONIC-BOOM GENERATION AND PROPAGATION


For a thorough overview of the details of the standard theory of sonic-boom propagation, the
reader is referred to Plotkin (1989), which contains, at a high level, an understandable description of
all the necessary concepts, some details of the implementation of sonic-boom-propagation codes,
and the considerations that must be accounted for in unsteady maneuvers (and focused booms)

www.annualreviews.org MDO for Sonic-Boom Minimization 509


FL44CH21-Alonso ARI 21 November 2011 13:0

a Aerodynamic b Acoustic
U U(t)
M > 1 M(t)

Ray tube

Shock wave
p
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

p
by Brown University on 12/14/12. For personal use only.

Figure 3
Sonic-boom propagation from the (a) aerodynamic (reference frame attached to the aircraft) and
(b) acoustics points of view (reference frame sitting on the ground).

and with atmospheric turbulence effects, as well as a brief section on the necessary modications
to account for nite rise times across shock waves when computing any loudness measure.
In this section, however, we focus on describing the fundamentals of classical sonic-boom-
propagation theory and point both to some of its shortcomings and to approaches to improve the
quality of the predictions. Figure 3 depicts two approaches to understanding the phenomenon of
sonic-boom propagation. In Figure 3a, we adopt a reference frame that is xed on the airplane
as it is typical in applied aerodynamics. During the cruise condition when the aircraft is ying
at constant altitude, velocity, and attitude, this reference frame is inertial, and we restrict our
comments in this article to such a situation. In this reference frame, the aircraft sees a supersonic
free stream coming toward it at a Mach number M that creates pressure disturbances as it
ows around the aircraft surfaces. These disturbances are conned to a region of space (the
region of inuence), a downstream-facing Mach cone, that is reachable by the nite speed of
propagation of the disturbances, the local speed of sound, a (see Figure 4). Figure 3b illustrates
the point of view that would be more traditionally associated with acoustics: We observe the
aircraft from an inertial reference frame that is xed on the ground. The aircraft moves with a
(possibly nonconstant) velocity U (t) and, at any instant in time, produces pressure disturbances
that travel away from the aircraft in a direction perpendicular to the acoustic wave fronts. We
adopt the traditional concept of a ray tube delimiting the region of space that determines the path
that a sound pressure disturbance will follow. The actual path and area distribution of this ray tube
will depend on many factors, including the altitude and speed of the aircraft but, more importantly,
the temperature/speed of sound distribution in the atmosphere that causes the refraction of the
ray tubes, and the presence of winds and atmospheric turbulence (Yamashita & Obayashi 2009).
In some sense, the two views of the problem are the same if we consider that the product of the
time t (since a pressure disturbance was created at the aircraft) and the aircraft speed U plays the

510 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

Top view

Flight
track

Mach cone

Side view

Boom
carpet
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

Figure 4
by Brown University on 12/14/12. For personal use only.

Three-dimensional view of the region of inuence from a supersonic aircraft. The intersection of the ray
tubes within the Mach cone that reach the ground forms the supersonic-boom carpet.

role of the x coordinate in theaerodynamicists point of view. In all subsequent formulae, is the
Prandtl-Glauert factor, = M 2 1, = C p is the ratio of specic heats for air (usually taken
Cv
to be 1.4), and s is the distance that a disturbance has traveled along a ray tube.
The theory of sonic-boom propagation describes the propagation of the pressure disturbances
created by the aircraft along the ray tubes until they reach the ground as the sonic-boom ground
signature. If one can create the paths that the ray tubes will follow (through suitable integration
across an atmosphere with changing temperature and therefore a changing speed of sound), the
solution of the sonic-boom-propagation problem reduces to the nonlinear wave propagation of a
complex signature along a ray tube of changing (but known) cross-sectional area. The portions of
the acoustic signature that reach the ground make up the boom carpet, which can be approximately
30 miles in width. (The refraction of acoustic rays in the atmosphere make it possible for only
a range of azimuthal disturbances to reach the ground: Typically the cutoff angle is around 55
from the ight track.) A secondary boom carpet can exist beyond a lateral zone of silence and is
created by rays that move upward from the aircraft into the upper atmosphere and are refracted
downward until they intersect the ground.
In general, it is customary to divide the problem into two parts. First, we use aerodynamic
theories to obtain the initial conditions of the problem, the source pressure waveform, and then
we propagate this waveform/signature along the ray tubes until it reaches the desired altitude,
typically the ground. In theory, the processes of generation and propagation can be described by
the same set of governing equations and the entire problem could be solved in one step. However,
there are reasons why this is not typically done.
The traditional approach to generating the initial aircraft pressure disturbance has been to
use classical linearized supersonic aerodynamics. Following Seebass (1969b), Ashley & Landahl
(1965), and Jones (1953), and using what is referred to as Hayess method, one can show that,
far enough from the aircraft (at a radial distance r  x r), but still sufciently close to it
(e.g., between 2 and 10 body lengths beneath the airplane), the pressure disturbance at any point
(x, r, ) comprises two contributions. The rst one comes from the cross-sectional area of the
aircraft when cut through a plane that is tangent to the upstream-facing Mach cone at (x, r, ) and

www.annualreviews.org MDO for Sonic-Boom Minimization 511


FL44CH21-Alonso ARI 21 November 2011 13:0

perpendicular to the corresponding (r, ) plane and is projected onto a plane perpendicular to the
free stream. This part of the signature is referred to as the volume-dependent portion. The second
contribution comes from the component of the force created by the surface pressure distribution
along the same cross section, perpendicular to the free stream and contained in the = const
plane. In other words, p(x, r; ) = p V (x, r; ) + p L (x, r; ). Relying on the mathematical
proof presented by Ashley & Landahl (1965), one can state that this pressure disturbance created
by the actual aircraft can also be thought of as being produced, for any particular value of , by
the equivalent area distribution, A(x) = AV (x) + AL (x). This equivalent area distribution is a
solid body of revolution that, for all practical purposes, can be thought of as producing the correct
pressure disturbance. Depending on the azimuthal location at which the pressure disturbance is
created, the equivalent body of revolution will be different. In other words, the pressure disturbance
depends on the second streamwise derivative of the equivalent area distribution as
 xr
p(x, r; ) 1 A ( ; )
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

  d. (1)
p M 2
2 2r 0 (x r)
It is customary to express this pressure disturbance in terms of the Whitham F-function,
 x
1 A ( )
by Brown University on 12/14/12. For personal use only.

F (x; ) = d, (2)
2 0 x
and Equation 1 is often presented as
M
2
F (x r; )
p(x r, r; ) = p  . (3)
2r
Given that, according to Equation 1, the pressure disturbance is the Abel transform of the second
derivative of the equivalent area distribution, it can easily be seen that the equivalent area distri-
bution, A(x), can be recovered as an inverse Abel transform of the pressure disturbance in the near
eld, which nowadays can be obtained by nonlinear computational uid dynamics (CFD) methods.
The equivalent area distribution has some properties that are worth mentioning. First, the area
distribution due to volume goes to zero at the front and back of AV (x) because the cutting planes
will eventually fail to intersect the aircraft. The distribution of AV (x) itself will increase or decrease
as more or less of the volume contained by the different elements of the aircraft (e.g., fuselage,
wing, empennage, nacelles) is intersected. Second, as the rear of the aircraft is approached, the
lift-dependent portion of the equivalent area AL (x; ) L Ucos2 , where L is the total lift of the

aircraft that, in the cruise condition, is equal to its weight, W. In other words, in supersonic ow
and for lifting bodies, the base (also called the caliber as the equivalent area distribution resembles
the shape of a bullet) of A(x) is nite.
However, this use of classical linearized supersonic aerodynamics, although powerful and in-
sightful, can suffer from signicant accuracy degradation because it assumes that the ow eld is
linear between the aircraft and the location underneath the vehicle at which the pressure signature
is computed. This assumption is typically violated as the strength of some shocks and expansions
(particularly those created by the wing and the inlet/nozzle of the propulsion system) can hardly be
considered a small perturbation over the free stream. This realization, together with the advent of
CFD for complex congurations, solution-adaptive methods, and powerful computing platforms,
has enabled the computation of the near-eld pressure signature by solving the Euler or Reynolds-
averaged Navier-Stokes equations of the ow and by extracting the pressure signature from the
computational mesh itself (see Campbell et al. 2008, Choi et al. 2009, Wintzer et al. 2010). As seen
in Figure 5 and explained by Choi et al. (2009), this method of computation presents a signicant
challenge: The initial pressure signature must be obtained sufciently far enough from the aircraft

512 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

a b

0.04
c
0.02
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

dp/p
0.02 No adaptation
1, uniform adaptation
2, uniform adaptation
0.04
by Brown University on 12/14/12. For personal use only.

3, solution adaptation
4,5, solution adaptation
1.4 1.8 10 15 20 25 30 35 40 45 50 55 60
Mach number x, distance from nose
of fuselage (inches)

Figure 5
Schematic for the extraction of a near-eld sonic-boom signature from a CFD computation, (a) symmetry-plane pressure distribution,
in which the grid has been adaptively rened (b) to achieve high accuracy. Actual extracted sample signatures show the inuence of
rening the mesh: Signatures for the nest meshes capture the discontinuities with the highest accuracy (c).

so that strong nonlinear effects have died down, resulting in CFD meshes with very large numbers
of elements for which the inherent dissipation in the numerical schemes does not articially de-
crease the strength of the near-eld signature and therefore the measure of sonic-boom strength
on the ground. However, the assumptions in typical sonic-boom-propagation schemes are not
necessarily consistent with the extraction of an initial pressure signature from CFD that is close to
the body of the aircraft (as is typically required to minimize the size of the computational mesh).
The issue is one of near-eld matching: How can we handle the cross-ow inherent in CFD
solutions that are taken only a short distance away from the aircraft itself ? Multipole expansions
to handle these issues have been proposed and used by Plotkin (2002) and represent the current
state of the art.
Once a near-eld sonic-boom signature has been obtained, it must be propagated to the ground
using the ideas and concepts described in the rst part of this section. As mentioned above, although
one could use the same governing equations that are utilized to compute the near-eld pressure
signature to propagate the sonic boom all the way to the ground, there are signicant problems with
both the classical linearized approach and a CFD-based one. With the classical linearized equations
of supersonics, the sonic boom would propagate all the way to the ground along characteristic lines
in the ow, which, in an isothermal atmosphere, would be straight lines inclined at the Mach angle
to the free stream. The shape of the near-eld pressure would be preserved, and the magnitude
of the pressure disturbances would only decay according to the rate of geometric expansion of
the area of the ray tubes. Both these conclusions are known to be incorrect: Nonlinear effects (wave
steepening and shock weakening) are critical to matching the experimentally observed behavior.

www.annualreviews.org MDO for Sonic-Boom Minimization 513


FL44CH21-Alonso ARI 21 November 2011 13:0

The alternative would be to create a sufciently ne computational grid all the way from the
aircraft (typically cruising at 60,000 ft) to the ground plane. This is also obviously unachievable
with nite computational resources and has not been pursued unless the studies are only interested
in obtaining the pressure signature at relatively close distances from the aircraft (for validation
with wind-tunnel experiments or ight test data obtained with two separate aircraft), as was done
for the SSBD.
The most common approach is to include rst-order corrections in the linearized equations
of supersonic ow and to propagate the signature along the ray tubes described above. The key
corrections introduced into the signature are (a) allowing for the local speed of propagation of any
point in the signature (along a ray tube) to be dependent on the magnitude of the disturbance itself
(essentially using a form of Burgers equation), (b) allowing for the formation of new shocks due
to nonlinear wave steepening (while avoiding multivalued solutions), (c) accounting for the proper
weakening of the shocks in the signature as they propagate along the ray tube, and (d ) accounting
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

for the correct changes in the area of the ray tube due to temperature gradients and stratied
winds in the atmosphere. The rst computer program to propagate sonic-boom signatures was
due to Hayes et al. (1969), but shortly after Thomas (1972) parameterized the problem in a sim-
pler form that led to improvements in robustness by eliminating the need for nding shocks in
by Brown University on 12/14/12. For personal use only.

the signature and using area-balancing conditions to locate the shocks themselves. Known as the
Thomas algorithm in the sonic-boom community, it has been the source of many improvements
in the series of PCBOOM programs by Plotkin et al. and remains the most commonly used ap-
proach today. For example, Wintzer et al. (2010) have applied this same procedure to the prediction
of the sonic-boom signatures of a number of test cases of increasing complexity, including the
DARPA/NASA/Northrop-Grumman SSBD shown in Figure 6. The agreement with the near-
eld pressure distributions obtained with a second supersonic aircraft ying below the SSBD is
quite good. However, although still relatively good, discrepancies exist in the comparisons of the
extrapolated ground-boom signatures and the experimental data collected during this program
(see Pawlowski et al. 2005). The sources of the discrepancies are not clear, but one can specu-
late that the differences result from (a) approximations made in the modeling of the propagation

a b c

0.050

0.025
p/p

0
h = 80 ft
0.025

0.050
0 200 400 600 800 1,000
s (inches)

Cp

Figure 6
Sonic-boom predictions for the DARPA/NASA/NG SSBD aircraft. (a) View of the adapted mesh on the symmetry plane. (b) Cp
contours on the symmetry plane with a modied forebody to create a shaped sonic-boom distribution. (c) Normalized pressure
disturbance along a line approximately two body lengths (80 ft) below the aircraft. Figure taken from Wintzer et al. (2010).

514 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

physics (inviscid, quasi-two-dimensional corrections to potential ow theory), (b) the effects of


turbulence in the atmosphere, and (c) the proper accounting of winds and other uncertainties in
the atmospheric properties.
Recent efforts in propagation methods have focused on improving the physical models involved
in the propagation schemes. Two such efforts are notable. Rallabhandi (2011) recently used an aug-
mented Burgers equation within the context of ray-tracing/geometrical acoustics to propagate the
source signatures to the ground while accounting for viscous effects that lead to nonzero thickness
shock discontinuities. Although the treatment of the effects of turbulence, winds, and atmospheric
uncertainties is the same as in more traditional methods, a signicant improvement in the state of
the art is that the shocks predicted at the ground have the physically correct thicknesses dictated
by molecular relaxation: Shocks no longer need to be articially thickened (with correlation-based
nite rise times) as a postprocessing step prior to computations of perceived loudness or structural
impact. Additional improvements in such methodologies for boom propagation can help reduce
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

discrepancies with ight-test data. Ozcer (2007) also attempted to improve on the state of the art
by solving the propagation of the signature using the full potential equation in the region between
the near eld and the ground. He used grid-adaptation and shock-tting techniques to obtain
a high-quality solution at the ground that is as close as possible to the true solution of the full
by Brown University on 12/14/12. For personal use only.

potential equation. As in the previous method, Ozcers work approaches some of the shortcomings
of traditional methodologies, but not all, leading to improved predictive capabilities but not to a
complete solution of the physics-based propagation of sonic booms.
Further improvements in our ability to propagate sonic booms with a high degree of accuracy
will result from simultaneously overcoming the shortcomings of traditional procedures. This
implies both using physically accurate models and accounting for the variability in the atmosphere,
including atmospheric turbulence effects. The challenge of a full simulation is daunting, but we
believe that the most logical way to proceed is to develop computational procedures with physically
accurate models (such as that of Rallabhandi) that account for the effects of turbulence, winds, and
atmospheric uncertainties by establishing computational bounds on the variations of the ground-
boom signatures that would result. Given the current focus by various institutions worldwide in
developing environmentally viable supersonic aircraft, we hope that some effort will be devoted
to this topic.

3. THE ACOUSTIC THEORY OF SONIC-BOOM MINIMIZATION


Once a reasonably accurate understanding of sonic-boom phenomena and propagation was estab-
lished, attention quickly turned to the possibility of inuencing the near-eld signature so that, as
it evolved to the mid- and far eld, its impact could be reduced/minimized. The most signicant
early efforts constitute the theory of sonic-boom minimization started by Jones (1961) and later
enhanced and given its well-known form by Seebass (1969b) with the help of George and Darden
(Darden 1977, George & Seebass 1971, Seebass & George 1973).
The original JSGD theory relies on a particular parameterization of the near-eld pressure
distribution and F-function (and therefore the resulting ground-boom signature) that allowed for
analytic optimization of the boom signature. This assumed form of the F-function was custom
designed to produce a family of signatures on the ground that were known a priori to be superior
based on certain boom metrics known at the time (maximum overpressure, signature impulse,
and initial pressure rise). In the classical theory of boom minimization, these metrics were used
as surrogates for outdoor and indoor perceived loudness/annoyance. Although our knowledge of
appropriate metrics has improved tremendously since then (see Coulouvrat 2009, for example), the
results produced by Seebass still retain much of their value. For these reasons, the parameterization

www.annualreviews.org MDO for Sonic-Boom Minimization 515


FL44CH21-Alonso ARI 21 November 2011 13:0

F
B
H
C

yf T


Figure 7
Sketch of general Jones-Seebass-George-Darden source signature parameterization, with key points noted.
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

includes at least one expansion shock (or pressure-drop discontinuity) at roughly the point at which
the trailing edge of the wing meets the fuselage. This parameterization is sketched in Figure 7
and is dened by
by Brown University on 12/14/12. For personal use only.

F (x) = 2H x/y f , 0 x y f /2,


F (x) = C(2x/y f 1) H (2x/y f 2), y f /2 x y f ,
(4)
F (x) = B(x y f ) + C, y f x ,
F (x) = B(x x f ) D, x L.

Holding the overall length of the aircraft constant, six variables remain to form the design vector:

x = {H , C, B, D, y f , }T . (5)

Using this piecewise-linear F-function parameterization, the lift constraint, and the free-stream
conditions, one can propagate the signal above analytically through an isothermal atmosphere
(Darden 1975) to obtain a given value of the boom-intensity measure of interest. As suggested
above, some early boom-strength metrics included the maximum pressure in the ground signal
(or overpressure), its maximum initial pressure rise (the strength of the leading shock), and the
maximum pressure impulse (positive area under the ground signal), which is a reasonable rst
step toward quantifying the impact of sonic booms on man-made structures. Figure 2 shows two
examples of these families of optimum solutions obtained using the JSGD F-function. For a full
presentation of the optimum families and predictions from the theory, the reader is referred to
the early works of Seebass and George (Seebass 1973, Seebass & George 1974) and the work
of Darden (1977). Many qualitative results found through these classical optimizations form the
basis of supersonic aircraft design principles still used in practice. In general, the minimum boom
strengths attainable through these simple boom-strength metrics increase with increasing ight
Mach number, increase with aircraft weight (under the assumption it is equal to lift), decrease
with aircraft length, and are largely insensitive to altitude (above approximately 50,000 ft). A nal
and important result of classical theory is the trade-off between nose bluntness (and hence drag)
and boom strength. Somewhat counterintuitively, a stronger shock at the source results in a more
rapid decay of the shock strength as it propagates through the atmosphere, resulting in a weaker
boom on the ground. This design trade has been widely studied since, and recently through
multiobjective optimization (Choi et al. 2008).
Although most work initially focused on minimizing some of the metrics mentioned above,
more realistic and robust metrics for boom noise, such as perceived loudness (PLdB) and

516 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

110 ft 4 inches

262 ft
Figure 8
Three-view drawing of the Boeing/NASA N+3 vehicle concept incorporating a canard to match the target
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

F-function/equivalent area distribution and closely integrated nacelles on the upper surface of the
conguration. Figure taken from Welge et al. (2010).

others (Makino & Kroo 2006), have been used for boom optimization and take into account
by Brown University on 12/14/12. For personal use only.

the realistic structure of shocks (most notably the nite rise time) and the reaction of the hu-
man ear (frequency content). All these advancements were built on the classical JSGD theory
to yield a number of current low-boom designs through analysis and, in some cases, ight tests.
These include the DARPA/NASA SSBD in 2003 (Pawlowski et al. 2005), which included some
1,300 measurements of sonic-boom signatures and a related follow-on program, the Quiet Spike
(NASA-Gulfstream collaboration, 2006) (Cowart & Grindle 2008, Howe et al. 2008). Addition-
ally, supersonic business-jet concepts have been studied by Gulfstream (Henne 2003), Boeing
(Welge et al. 2010), and Lockheed-Martin (Morgenstern et al. 2010). An example of the Boeing
design for a NASA N+3 concept is shown in Figure 8.
Fundamental to many current low-boom concepts is a sting (or nose extension) that extends
the apparent length of the vehicle and has a benecial effect on boom amplitudes. In particular, the
Quiet Spike concept of Gulfstream (Howe et al. 2008) consists of alternating inclined and parallel
regions, creating a sequence of weak shocks as compared with a standard nose geometry (see
Figure 9). As a signal propagates, shocks will weaken and expansion regions will spread (decrease

Shocks
Expansions

Figure 9
Schematic diagram and pressure plot of a sting shaped for multiple shocks. As the signal propagates, the
intermediate expansion regions prevent coalescence of the individual shocks into a single, stronger shock,
resulting in a lower-boom signal.

www.annualreviews.org MDO for Sonic-Boom Minimization 517


FL44CH21-Alonso ARI 21 November 2011 13:0

in slope and increase the distance between the end points). Hence placing expansion regions
between smaller shocks will prevent the shocks from coalescing into a single, stronger shock and
ultimately result in a considerably weaker boom on the ground (a more modern tailored signature).
This process can be generalized to three-dimensional geometries and entire aircraft, shaping the
forward portion of the vehicle to achieve a desired trade-off between boom strength and drag
while maintaining the overall lift of the vehicle. This MDO approach consisting of a fully three-
dimensional design that accounts for aerodynamic and boom considerations represents the current
frontier in numerical sonic-boom research.
Having briey reviewed the classical theory of sonic-boom minimization above, we now discuss
its limitations in the design of modern supersonic aircraft. The classical theory still remains an
extremely valuable tool for the aircraft designer, but modications offered by a more modern
view can be used to update the tools available to aircraft designers looking at current-day low-
boom aircraft designs. Here we cite four specic limitations for further study. The limitations of
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

linearized supersonic ow theory, discussed above for near-eld pressure signatures, are noted
here but not discussed further as, if needed, CFD solutions can be used in place of F-function-
based aerodynamics. The F-function method is an adequate approximation for reasonably slender
vehicles without a blunt nose at moderate supersonic Mach numbers. Hypersonic, short, and
by Brown University on 12/14/12. For personal use only.

blunter vehicles require an alternative method for near-eld pressure prediction.

3.1. Aircraft Parameterization


As discussed above, F-functions have traditionally been parameterized with a handful (four to
ve) of parameters in specic ways amenable to (a) generalized a priori aircraft congurations and
(b) analytic optimizations. When broadening the scope of the approach to be based on numerical
optimization, not analytic formulations, one can also generalize the parameterization of the aircraft
near-eld pressure distribution to higher-dimensional spaces with the hope that optimum solutions
not allowed by the classical theory can be captured. For example, Colonno & Alonso (2010) used
a parameterization of the F-function including 13 parameters. The reason for this signicant
expansion in the number of parameters (compared with classical theory) is that over the past
decade there has been evidence that ground-boom signatures with multiple shocks (created by
slope discontinuities in the surface of the aircraft, such as the Gulfstream/NASA Quiet Spike) can
be quiet and possibly more robust to small changes in the input parameters than the traditional
solutions offered by the classical theory.

3.2. Effects of Uncertainties and Robust Boom Signatures


Perhaps the most signicant limitation in the standard theory is the use of a purely deterministic
approach. Although the solutions found are optimum within the range of the design space in the
problem formulation, they are not necessarily robust. In other words, do small perturbations in the
optimum parameters lead to dramatic changes in the resulting ground-boom signatures? Because
of the natural variability in atmospheric properties, in the lift magnitude and distribution (both
because of changes in weight during an actual ight and because of different passenger and fuel
loading), and in the engine-operating parameters, deterministic optima are not likely to be well
suited to actual operating aircraft. To date, we are not aware of any formal work in the design
of robust sonic-boom signatures that have low loudness levels but are also resistant to variations
in many operating and environmental parameters. The application of uncertainty quantication
techniques is ripe for new developments on this topic.

518 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

3.3. Aerodynamic Performance Constraints


A given F-function parameterization determines, within the delity of linearized supersonic the-
ory, the lift and drag properties of the aircraft as well as the near-eld pressure signature. Hence
the F-function parameters cannot be independently varied in an optimization; constraints must
be placed on aerodynamic performance to ensure a feasible aircraft design. The cruise lift of an
aircraft must be constrained to its weight (which varies substantially over the course of a ight).
In addition, the wave drag of the aircraft can be evaluated (although not including viscous drag)
within the connes of linear theory and, if desired, constrained to address the well-known trade-
off between nose bluntness and boom strength and to reect higher-level trade-offs in the overall
performance of the vehicle. Studies that include aerodynamic constraints in the optimization
framework are needed.
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

3.4. Metrics for Outdoor Sonic-Boom Perception


The analytic methods used in the classical theory and the lack of accepted loudness metrics (at
the time) prevented earlier work from using outdoor loudness metrics that can be correlated with
the annoyance to the population. Additional metrics are now available and can be used as objective
by Brown University on 12/14/12. For personal use only.

functions in optimizations. This is primarily as a result of the available computational power and the
use of a genetic-algorithm-based optimizer for which smooth gradients of the objective function
are not required. As outdoor and indoor boom-strength metrics relevant to human hearing evolve,
they can easily be incorporated into the framework discussed here. In addition, the standard theory
is based on isentropic small-amplitude wave propagation, and shocks are treated as zero-thickness
discontinuities. Real shocks have a nite rise time that is important to the high-frequency portion
of the boom spectrum and that must be accounted for within the optimization to obtain accurate
ground signatures and boom strengths (especially when closely spaced shocks are present in the
nal signal).

4. MULTIDISCIPLINARY OPTIMIZATION AND LOW SONIC BOOM


If we assume that (a) a near-eld pressure disturbance can be accurately propagated to the ground
(typically in the mid-eld of the signature), (b) a proper metric of boom loudness/annoyance
can be dened, (c) the characteristics of signatures that result in desirable values of the cho-
sen metric are known, and (d ) clear mission requirements/constraints have been provided (e.g.,
payload, range, subsonic and supersonic cruise Mach numbers, takeoff and landing eld length,
emissions, noise, boom loudness), then in principle it should be possible to attempt to design
a supersonic aircraft with low boom that meets all the other requirements/constraints in the
safest, most-efcient manner. This statement assumes that there is an actual aircraft congura-
tion that meets all the design constraints, which may not be the case if the constraints are too
stringent. When the design problem is worded in such a way, one is immediately drawn to the
theory of optimization, involving the interaction of multiple disciplines that dictate the overall
performance of the resulting aircraft. The eld of MDO is the body of knowledge that facilitates
the successful completion of such optimal design problems by carefully ensuring that the com-
ponent disciplines and their interactions are properly accounted for. For them to be realizable,
supersonic aircraft designs typically exhibit close integration of the component disciplines. As is
evident from some recently proposed congurations, the aerodynamics and propulsion systems
are intimately coupled to the point that they are barely distinguishable. In addition, structures,
stability and control, mission proles, and sonic boom also have signicant interactions among

www.annualreviews.org MDO for Sonic-Boom Minimization 519


FL44CH21-Alonso ARI 21 November 2011 13:0

them. The problem of designing such closely integrated aircraft is unfortunately easier said than
done.
To illustrate the challenges of using MDO to achieve a low-boom design, we use the design
requirements of the NASA Supersonics Project for the N+3 generation of supersonic aircraft.
NASA is intent on overcoming technology barriers so that an aircraft of this kind could be available
in the 20302035 time frame. More details about the NASA Supersonics Project can be found
elsewhere (Fund. Aeronaut. Prog. 2008). The objectives of the N+3 effort are to design an efcient
multi-Mach-capable aircraft that can y 100200 passengers with a minimum range of 4,000 nmi,
at cruise speeds between Mach 1.3 and 2.0. (Higher speeds may be used when ying over water
and the restrictions on sonic boom are not as severe, whereas lower speeds may be necessary
to meet the sonic-boom requirements when ying over land.) The sonic boom must be such
that the outdoor perceived loudness of the aircraft is between 65 and 70 PLdB for low-boom
ight, and between 75 and 80 PLdB for unrestricted ight. In addition, airport noise must be 20
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

30 EPNdB (cumulative) below stage-3 requirements, cruise emissions (grams of NOx/Kg fuel)
must be less than ve, efforts must be made for particulate and water vapor mitigation, and the
fuel efciency (in passenger miles/pounds of fuel) must be between 3.5 and 4.5. Under contract
to NASA, Lockheed-Martin pursued a conceptual-level study of various candidate congurations
by Brown University on 12/14/12. For personal use only.

that resulted in a preferred design concept (Morgenstern et al. 2010).


Presently, a fully automated procedure to produce an aircraft design that meets all of its con-
straints and that is optimal in some designer-specied ways is beyond the reach of MDO capabil-
ities. However, when a starting design concept is appropriately created and parameterized (using
information from the physics of each discipline and prior design experience), many techniques
are available to ensure that the design (a) is feasible (it meets all the constraints) and (b) is optimal
in the sense that it is better than any other concept in some ways that are specied by the design
team. Figure 10 shows a number of concepts that were explored in the NASA/Lockheed-Martin

Family of configurations

Initial configuration definition: inverted-V Oblique flying wing configuration

Engine over wing configuration Twin fuselage configuration

Mother/ Blue sky


daughter session

T-tail configuration Mother/daughter, Blue sky alternatives

Figure 10
A variety of initial concepts proposed in a NASA N+3 study pursued by Lockheed-Martin. The inverted-V
conguration was selected as the design with the most potential to meet all requirements while maximizing
performance. Figure taken from Morgenstern et al. (2010).

520 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

design study. The variety of initial congurations is a direct representation of the impossibility,
before the design process, of foreseeing the key characteristics of the optimal design. Viable ap-
proaches include inverse design procedures in which a target signature in the near eld is dened
and changes in the aircraft geometry to match that signature are sought (Li et al. 2008).
MDO is often accomplished by casting the design-optimization problem into the form of a
nonlinear program (Gill et al. 1982). Formally, we choose to nd the value of a vector of design
parameters x that minimizes a cost/objective function of interest f ( x ), with x R N , such that
the linear or nonlinear constraints gi ( x ) 0, i = 1, . . . , M are satised. The dimensionality of
the design parameter vector x can be rather large, as can the number of constraints. In most
aerodynamic shape-optimization problems, however, typically N  M. To evolve the initial
design into an optimal one (that satises the Kuhn-Tucker optimality conditions), optimizers
typically use the repeated evaluation of the objective function and constraints and sometimes their
sensitivities with respect to the design parameters, as is the case in gradient-based optimization
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

techniques. For example, Choi et al. (2008) used such an approach to carry out optimal supersonic
designs using a variety of different problem formulations. The reader is reminded that, regardless of
whether sonic boom is considered a constraint (which must be met or exceeded) or a cost function
(which must be minimized), every evaluation of the cost function and/or constraints requires
by Brown University on 12/14/12. For personal use only.

the analysis of the near-eld signature, the propagation of this signature to the far eld, and the
computation of some loudness/annoyance metric resulting from the ground-boom signature. The
repeated evaluation of these functions leads to signicant computational costs (akin to adding more
dimensions into a problem, such as when one goes from predicting steady-state ows to unsteady
ones) that are often mitigated through the use of surrogate models or response surfaces. These
can be precomputed by judiciously chosen analyses spanning the entire design space.
Specically in supersonic design, however, there are several practical barriers that are being
currently researched. The problems that must be overcome typically fall under the following three
broad categories.

4.1. Single-Objective or Multiobjective Design Procedures


The NASA N+3 problem description provided at the beginning of this section does not nec-
essarily resemble an optimization problem because there are no specic functions that must be
maximized/minimized. Instead, the problem is formulated as a constraint-feasible problem: Can
we nd a design that at least meets the constraints of the problem? However, if more than one
feasible design exists, it behooves us to choose among those designs the one that is best in some
measure. In some senses, combinations of constraints and objective functions could be applied to
the N+3 environmental and performance goals to create a proper MDO problem. For example,
one may be required to meet constraints involving the Mach number, emissions, fuel efciency,
payload, and range, while minimizing the sonic-boom loudness over land. Alternatively, one may
be asked to ensure that the sonic-boom loudness be no higher than a given value and to accom-
plish that (together with all the other constraints) with an aircraft with the lowest maximum takeoff
gross weight, which is often positively correlated with the cost of the vehicle. These are examples
of single-objective nonlinearly constrained optimization.
However, by choosing values of the constraints and single-mindedly focusing on an individual
objective function to optimize, we might, by construction, miss the opportunity to nd a design
that is better. In other words, good designs may be precluded by the problem formulation chosen.
In contrast, if we formulated the problem with the necessary constraints to be satised and then
we asked the optimizer to pursue the minimization of multiple objectives simultaneously (e.g.,
minimizing both the sonic-boom loudness and the coefcient of drag, CD , of the airplane at a

www.annualreviews.org MDO for Sonic-Boom Minimization 521


FL44CH21-Alonso ARI 21 November 2011 13:0

constant cruise lift coefcient, CL ), we would arrive at a family of solutions that, while satisfying
all constraints in the problem, would provide the designer with reasonable trade-offs between the
attainable values of the two objective functions. This Pareto-optimal family of solutions represents
multiple designs that are not strictly worse than any other solution in both objectives simultane-
ously. Although the costs of arriving at such families of solutions may be excessively high, mul-
tiobjective optimization has emerged as a desirable approach particularly for exploratory studies
of supersonic low-boom jets. Choi et al. (2004) explored this approach and, in conjunction with
surrogate models of both the aerodynamic performance and the loudness of the sonic boom at
the ground, and genetic algorithms, were able to compute Pareto fronts that shed light on the
conguration trade-offs between boom and performance.
Such approaches are quickly becoming viable, particularly in conjunction with multidelity
techniques that combine low- and high-delity analysis tools to produce results of high delity
at much lower computational costs. Additional research issues remain, for example, ensuring
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

that sufcient (and even) coverage of the Pareto front results from the computations, as well
as interpreting and deriving design information from multidimensional Pareto fronts that may
involve multiple objectives.
by Brown University on 12/14/12. For personal use only.

4.2. Topology and Continuity of the Design Space


A signicant challenge presented by the design of low-boom aircraft is that, whereas typical aircraft
design evolves within design spaces that are relatively well behaved (cost and constraint functions
are relatively smooth and often continuous functions of the design parameters), many boom-
related metrics are not well behaved (Chung & Alonso 2002). In fact, they can be discontinuous
close to the optimum, and often can be relatively noisy because of the physics (small changes in
design parameters can result in the coalescence of shocks in multishock signatures), the choice of
cost functions and constraints, and the actual software implementation. All these characteristics
of low-boom design spaces make the convergence of the optimization process challenging at best.
Approaches to regularize the design space, to improve the robustness/noisiness of the predictions,
to produce sensitivity values that are accurate, and to handle uncertainties inherent to the variability
in the design space (which are critical for robust design strategies) are required and are seeing
ongoing research efforts at this time.

4.3. Design Parameterization, Boom Objective Formulation, and Efficient


Design/Search Algorithms
In this subsection, we lump a number of other research challenges in the MDO of low-boom
supersonic jets and simply enumerate them with a short description and a number of references
for further reading. The parameterization of the design is critical: It must be both sufciently
low dimensional and, at the same time, provide some certainty that the actual optimal solution
will be contained in the parameter space. Unfortunately, these two requirements often conict:
To construct a complete basis of the design space, one may have to use a very large number of
design parameters. In addition, a guarantee that the chosen parameterization contains the true
optimum of the system is a difcult task: One cannot easily draw conclusions on that which is not
observed during the design process. Experience has shown that parameterizations of complete
supersonic aircraft with as few as 50 parameters and as many as 500 (Reuther et al. 1999) can
be successfully used for the drag minimization of supersonic vehicles. However, there is little
direct evidence on the size of the parameterization that is required for supersonic design problems
involving sonic-boom minimization (Choi et al. 2008). Although some authors have used as few

522 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

as 10 design parameters and others have pushed the boundary with closer to 30, it is possible that
higher-dimensional discretizations may be required, especially if robust multishock signatures are
sought as the solution to the design problem.
The choice of actual objective function for sonic-boom minimization has a critical impact
on the ability to use optimization algorithms for this problem. Objective functions that lead to
discontinuous design spaces (such as the initial pressure rise, which can quickly change when two
shocks in a multishock signature coalesce) can render the design problem impossible. Recent work
(Makino & Kroo 2006) has attempted to address such difculties, but a better understanding of
the impact of new metrics (that address both indoor and outdoor boom impacts) is still needed.
Finally, optimization requires repetitive simulation of the system/aircraft with variations of the
design parameters. Algorithms for single- and multiobjective optimization that are more efcient
than those that exist today will still pace the rate of progress in the design of future low-sonic-boom
aircraft.
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

5. SUMMARY
This article presents some aspects of the historical accomplishments, current status, and future
by Brown University on 12/14/12. For personal use only.

research avenues required for the MDO of low-boom, supersonic aircraft. Classical linearized
theories have given way to CFD-based boom prediction and design techniques. MDO techniques
are currently being used to arrive at balanced designs that meet the quiet-boom requirements.
Using this entire body of knowledge, in the past 10 years researchers have seen the rapid develop-
ment of advanced concepts for supersonic aircraft that could one day carry signicant numbers of
passengers at supersonic speeds over land. The community is now beginning to see that a realistic
aircraft with a shaped signature that could permit supersonic ight over land is possible. Once
analytic and computational techniques have clearly shown that this is the case, the remaining re-
search task would be to demonstrate, through ight test and under various operating conditions,
that low-boom characteristics can be retained and that the resulting noise is almost imperceptible.

FUTURE ISSUES
1. Physics-based modeling of the propagation of sonic booms including atmospheric tur-
bulence and nite shock rise times is needed.
2. We need a new way of looking at the classical theory of sonic-boom minimization that
includes (a) high-dimensional parameterizations of the near-eld pressure signature,
(b) the quantication of the uncertainties present (and robust signature issues), and
(c) the consideration of more relevant sonic-boom metrics that incorporate all the knowl-
edge accumulated over the past 2030 years.
3. Flexible multidisciplinary optimization techniques are needed that can deal with com-
plex design spaces (discontinuous and noisy) that are likely to be found in low-boom
optimization.
4. Advanced experimental techniques are required to develop a database for the prediction
of shaped sonic booms.
5. Ultimately, researchers need to develop a ight test article that can be used as a research
platform for truly shaped sonic booms so that the full potential of low-boom aircraft can
be examined.

www.annualreviews.org MDO for Sonic-Boom Minimization 523


FL44CH21-Alonso ARI 21 November 2011 13:0

DISCLOSURE STATEMENT
Some of the authors work in the MDO of supersonic, low-boom aircraft is currently funded by
the NASA Supersonics Project.

ACKNOWLEDGMENTS
J.J.A. would like to acknowledge the support and contributions of NASA, the USAF, and DARPA
to many of the efforts discussed in this article. In addition, the contributions of former students
Hyoung S. Chung and Seongim Choi are gratefully acknowledged.

LITERATURE CITED
Alonso J, Kroo I, Jameson A. 2002. Advanced algorithms for design and optimization of quiet supersonic platforms.
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

Presented at AIAA Aerosp. Sci. Meet. Exhib., 40th, Reno, NV, AIAA Pap. 2002-0144
Aronstein D, Schueler K. 2005. Two supersonic business aircraft conceptual designs, with and without sonic
boom constraint. J. Aircr. 42:77586
Ashley H, Landahl M. 1965. Aerodynamics of Wings and Bodies. Reading, MA: Addison-Wesley
by Brown University on 12/14/12. For personal use only.

Campbell R, Carter M, Deere K, Waithe K. 2008. Efcient unstructured grid adaptation methods for sonic boom
prediction. Presented at AIAA Appl. Aerodyn. Conf., 26th, Honolulu, AIAA Pap. 2008-7327
Cheung S. 1994. Sonic boom softening of reference-H. Tech. Rep., NASA Langley Res. Cent., Orlando, FL
Cheung S, Edwards T, Lawrence S. 1992. Application of CFD to sonic boom near- and mid-eld prediction.
J. Aircr. 29:92026
Chiba K, Makino Y, Takatoya T. 2008. Evolutionary-based multidisciplinary design exploration for the silent
supersonic technology demonstrator wing. J. Aircr. 45:148194
Choi S, Alonso J, Kroo I, Wintzer M. 2008. Multidelity design optimization of low-boom supersonic jets.
J. Aircr. 45:10618
Choi S, Alonso JJ, van der Weide E. 2009. Numerical and mesh resolution requirements for accurate sonic
boom prediction. J. Aircr. 46:112639
Choi S, Chung HS, Alonso JJ. 2004. Design of low-boom supersonic business jet with evolutionary algorithms
using adaptive unstructured mesh. Presented at AIAA/ASME/ASCE/AHS/ASC Struct., Struct. Dyn. Mater.
Conf., 45th, Palm Springs, CA, AIAA Pap. 2004-1758
Chung HS, Alonso JJ. 2002. Design of a low-boom supersonic business jet using Cokriging approximation models.
Presented at AIAA/ISSMO Symp. Multidiscip. Anal. Optim., 9th, Atlanta, AIAA Pap. 2002-5598
Collard D. 1991. Concorde airframe design and development. SAE Trans. 100:262041
Colonno M, Alonso J. 2010. Sonic boom minimization revisited: the robustness of optimal low-boom designs. Presented
at AIAA/ISSMO Multidiscip. Anal. Optim. Conf., 13th, Fort Worth, TX, AIAA Pap. 2010-9364
Coulouvrat F. 2009. The challenges of dening an acceptable sonic boom overland. Presented at AIAA/CEAS
Aeroacoust. Conf. (30th AIAA Aeroacoust. Conf.), 15th, Miami, AIAA Pap. 2009-3384
Cowart R, Grindle T. 2008. An overview of the Gulfstream/NASA quiet spike ight test program. Presented at
AIAA Aerosp. Sci. Meet. Exhib., 46th, Reno, NV, AIAA Pap. 2008-123
Darden C. 1975. Minimization of sonic-boom parameters in real and isothermal atmospheres. Tech. Rep. TN
D-7842, NASA
Darden C. 1977. Sonic boom theory: its status in prediction and minimization. J. Aircr. 14:56976
Fund. Aeronaut. Prog. 2008. Fundamental aeronautics program overview. http:// www.aeronautics.nasa.gov/
fap/documents.html
George A, Seebass R. 1971. Sonic boom minimization including both front and rear shocks. AIAA Tech. Notes
9:209193
Gill PE, Murray W, Wright M. 1982. Practical Optimization. New York: Academic
Gold T. 1952. The double bang of supersonic aircraft. Nature 170:808
Hayes W. 1947. Linearized supersonic ow. North Am. Aviat. Rep. AL-222, Calif. Inst. Technol.

524 Alonso Colonno


FL44CH21-Alonso ARI 21 November 2011 13:0

Hayes WD, Haefeli RC, Kulsrud HE. 1969. Sonic boom propagation in a stratied atmosphere, with computer
program. Tech. Rep. NASA/CR 1299, NASA
Henne P. 2003. The case for small supersonic civil aircraft. Presented at AIAA/ICAS Int. Air Space Symp. Expo.,
Dayton, OH, AIAA Pap. 2003-2555
Henne P. 2005. Case for small supersonic civil aircraft. J. Aircr. 42:76574
Howe D, Waithe K, Haering E. 2008. Quiet spike near eld ight test pressure measurements with computational
uid dynamics comparisons. Presented at AIAA Aerosp. Sci. Meet. Exhib., 46th, Reno, NV, AIAA Pap. 2008-
128
Jones L. 1961. Lower bounds for sonic bangs. J. R. Aeronaut. Soc. 65:14
Jones R. 1953. Theory of wing-body drag at supersonic speeds. Tech. Rep. RM A53H18a, NACA
Jung T, Starkey R, Argrow B. 2011. Feasibility study of using a small-scale UAS for sonic boom minimization research.
Presented at AIAA Aerosp. Sci. Meet., 49th, Orlando, FL, AIAA Pap. 2011-1280
Li W, Shields E, Le D. 2008. Interactive inverse design optimization of fuselage shape for low-boom supersonic
concepts. J. Aircr. 45:138197
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

Makino Y, Kroo I. 2006. Robust objective functions for sonic-boom minimization. J. Aircr. 43:13016
McLean FE. 1965. Some nonasymptotic effects on the sonic boom of large aircraft. Tech. Rep. NASA/TN
D-2877, NASA
McLean FE. 1968. Conguration design for specied pressure signature characteristics. Tech. Rep. NASA/SP
180, NASA
by Brown University on 12/14/12. For personal use only.

Morgenstern J, Norstrud N, Stelmack M, Skoch C. 2010. Final report for the advanced concept studies for
supersonic commercial transports entering service in the 2030 to 2035 period, N+3 supersonic program.
Tech. Rep. NASA/CR 2010-216796, Lockheed Martin Aeronaut. Co.
Natl. Res. Council. 1997. U.S. supersonic commercial aircraft: assessing NASAs high speed research program.
Tech. Rep., Natl. Res. Council, Washington, DC
Ozcer I. 2007. Sonic boom prediction using Euler/full potential methodology. Presented at AIAA Aerosp. Sci. Meet.
Exhib., 45th, Reno, NV, AIAA Pap. 2007-369
Pawlowski J, Graham D, Boccadoro C, Coen P, Maglieri D. 2005. Origins and overview of the shaped sonic boom
demonstration program. Presented at AIAA Aerosp. Sci. Meet. Exhib., 43rd, Reno, NV, AIAA Pap. 2005-
0005
Plotkin K. 1989. Review of sonic boom theory. Presented at Aeroacoust. Conf., 12th, San Antonio, TX, AIAA
Pap. 89-1105
Plotkin K. 2002. Extrapolation of sonic boom signatures from CFD solutions. Presented at AIAA Aerosp. Sci. Meet.
Exhib., 40th, Reno, NV, AIAA Pap. 2002-922
Rallabhandi S. 2011. Advanced sonic boom prediction using augmented Burgers equation. Presented at 49th AIAA
Aerosp. Sci. Meet., Orlando, FL, AIAA Pap. 2011-1278
Reuther JJ, Jameson A, Alonso JJ, Rimlinger MJ, Saunders D. 1999. Constrained multipoint aerodynamic
shape optimization using an adjoint formulation and parallel computers, parts 1 and 2. AIAA J. Aircr.
36:5174
Seebass A. 1973. The design or operation of aircraft to minimize their sonic boom. Presented at 5th AIAA Aircr.
Des., Flight Test, Oper. Meet., St. Louis, AIAA Pap. 1973-817
Seebass A, Argrow B. 1998. Sonic boom minimization revisited. Presented at Theoret. Fluid Mech. Meet., 2nd,
Albuquerque, NM, AIAA Pap. 1998-2956
Seebass A, George A. 1973. Sonic boom reduction through aircraft design and operation. AIAA Aerosp. Sci. Meet.,
11th, Washington, DC, AIAA Pap. 1973-241
Seebass A, George A. 1974. Design and operation of aircraft to minimize their sonic boom. J. Aircr. 11:50917
Seebass R. 1969a. Minimum sonic boom shock strengths and overpressures. Nature 221:65153
Seebass R. 1969b. Sonic boom theory. J. Aircr. 6:17784
Seebass R, George A. 1972. Sonic-boom minimization. J. Acoust. Soc. Am. 51:68694
Tetzloff I, Crossley W. 2011. Evaluating market and environmental impacts of an N+3 supersonic aircraft. Pre-
sented at AIAA Aerosp. Sci. Meet. Exhib., 49th, Orlando, FL, AIAA Pap. 2011-463
Thomas C. 1972. Extrapolation of sonic boom pressure signatures by the waveform parameter method. Tech.
Rep. D-6832, NASA

www.annualreviews.org MDO for Sonic-Boom Minimization 525


FL44CH21-Alonso ARI 21 November 2011 13:0

Warren CHE. 1953. Noise from aircraft at supersonic speeds. Nature 171:21416
Welge H, Nelson C, Bonet J. 2010. Supersonic vehicle systems for the 2020 to 2035 timeframe. Presented at AIAA
Appl. Aerodyn. Conf., 28th, Chicago, AIAA Pap. 2010-4930
Whitham G. 1952. The ow pattern of a supersonic projectile. Commun. Pure Appl. Math. 5:30148
Wintzer M, Nemec M, Aftosmis M. 2010. Adjoint-based adaptive mesh renement for sonic boom prediction.
Presented at AIAA Appl. Aerodyn. Conf., 26th, Honolulu, AIAA Pap. 2008-6593
Yamashita H, Obayashi S. 2009. Sonic boom variability due to homogeneous atmospheric turbulence. J. Aircr.
46:188693
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org
by Brown University on 12/14/12. For personal use only.

526 Alonso Colonno


FL44-FrontMatter ARI 1 December 2011 22:45

Annual Review of
Fluid Mechanics

Contents Volume 44, 2012

Aeroacoustics of Musical Instruments


Benoit Fabre, Joel Gilbert, Avraham Hirschberg, and Xavier Pelorson p p p p p p p p p p p p p p p p p p p p 1
Cascades in Wall-Bounded Turbulence
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

Javier Jimenez p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p27


Large-Eddy-Simulation Tools for Multiphase Flows
Rodney O. Fox p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p47
by Brown University on 12/14/12. For personal use only.

Hydrodynamic Techniques to Enhance Membrane Filtration


Michel Y. Jaffrin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p77
Wake-Induced Oscillatory Paths of Bodies Freely Rising
or Falling in Fluids
Patricia Ern, Frederic Risso, David Fabre, and Jacques Magnaudet p p p p p p p p p p p p p p p p p p p p p p97
Flow and Transport in Regions with Aquatic Vegetation
Heidi M. Nepf p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 123
Electrorheological Fluids: Mechanisms, Dynamics,
and Microuidics Applications
Ping Sheng and Weijia Wen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 143
The Gyrokinetic Description of Microturbulence in Magnetized Plasmas
John A. Krommes p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 175
The Signicance of Simple Invariant Solutions in Turbulent Flows
Genta Kawahara, Markus Uhlmann, and Lennaert van Veen p p p p p p p p p p p p p p p p p p p p p p p p p p 203
Modern Challenges Facing Turbomachinery Aeroacoustics
Nigel Peake and Anthony B. Parry p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 227
Liquid Rope Coiling
Neil M. Ribe, Mehdi Habibi, and Daniel Bonn p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 249
Dynamics of the Tear Film
Richard J. Braun p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 267
Physics and Computation of Aero-Optics
Meng Wang, Ali Mani, and Stanislav Gordeyev p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 299

v
FL44-FrontMatter ARI 1 December 2011 22:45

Smoothed Particle Hydrodynamics and Its Diverse Applications


J.J. Monaghan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 323
Fluid Mechanics of the Eye
Jennifer H. Siggers and C. Ross Ethier p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 347
Fluid Mechanics of Planktonic Microorganisms
Jeffrey S. Guasto, Roberto Rusconi, and Roman Stocker p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 373
Nanoscale Electrokinetics and Microvortices: How Microhydrodynamics
Affects Nanouidic Ion Flux
Hsueh-Chia Chang, Gilad Yossifon, and Evgeny A. Demekhin p p p p p p p p p p p p p p p p p p p p p p p p p 401
Two-Dimensional Turbulence
Guido Boffetta and Robert E. Ecke p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 427
Annu. Rev. Fluid Mech. 2012.44:505-526. Downloaded from www.annualreviews.org

Vegetable Dynamicks: The Role of Water in Plant Movements


Jacques Dumais and Yoel Forterre p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 453
by Brown University on 12/14/12. For personal use only.

The Wind in the Willows: Flows in Forest Canopies in Complex Terrain


Stephen E. Belcher, Ian N. Harman, and John J. Finnigan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 479
Multidisciplinary Optimization with Applications
to Sonic-Boom Minimization
Juan J. Alonso and Michael R. Colonno p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 505
Direct Numerical Simulation on the Receptivity, Instability,
and Transition of Hypersonic Boundary Layers
Xiaolin Zhong and Xiaowen Wang p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 527
Air-Entrainment Mechanisms in Plunging Jets and Breaking Waves
Kenneth T. Kiger and James H. Duncan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 563

Indexes

Cumulative Index of Contributing Authors, Volumes 144 p p p p p p p p p p p p p p p p p p p p p p p p p p p p 597


Cumulative Index of Chapter Titles, Volumes 144 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 606

Errata

An online log of corrections to Annual Review of Fluid Mechanics articles may be found
at http://uid.annualreviews.org/errata.shtml

vi Contents

Potrebbero piacerti anche