personal homepages for the university of bath - stability and...

32
Annu. Rev. Fluid Mech. 2003. 35:413–40 doi: 10.1146/annurev.fluid.35.101101.161045 Copyright c 2003 by Annual Reviews. All rights reserved STABILITY AND TRANSITION OF THREE-DIMENSIONAL BOUNDARY LAYERS William S. Saric 1 , Helen L. Reed 1 , and Edward B. White 2 1 Mechanical and Aerospace Engineering, Arizona State University, Tempe, Arizona 85287-6106; email: [email protected], [email protected] 2 Mechanical and Aerospace Engineering, Case Western Reserve University, Cleveland, Ohio 44106-7222; email: [email protected] Key Words roughness, nonlinear, saturation, secondary instabilities, rotating disk, swept wing Abstract The recent progress in three-dimensional boundary-layer stability and transition is reviewed. The material focuses on the crossflow instability that leads to transition on swept wings and rotating disks. Following a brief overview of instability mechanisms and the crossflow problem, a summary of the important findings of the 1990s is given. 1. INTRODUCTION In low-disturbance environments such as flight, boundary-layer transition to tur- bulence generally occurs through the uninterrupted growth of linear instabilities. The initial conditions for these instabilities are introduced through the receptiv- ity process, which depends on a variety of factors (Saric et al. 2002). There is no shortage of publications in the field of boundary-layer stability and transi- tion. Comprehensive reviews for both two- and three-dimensional (3-D) flows are given by Arnal (1994), Mack (1984), and Reshotko (1994). Reed et al. (1996) give an up-to-date discussion of the effectiveness and limitations of linear the- ory in describing boundary-layer instabilities. The reader is referred to these re- ports for overviews of much of the early work in stability and transition. Crouch and coworkers (Crouch 1997, Crouch & Ng 2000, Crouch et al. 2001) and Herbert (1997b) have presented analyses of transition with specific applications to flight. Reed & Saric (1989) review the early theoretical and experimental work in 3-D boundary layers. Swept wings, rotating disks, axisymmetric bodies (rotating cones and spheres), corner flows, attachment-line instabilities, as well as the stability of flows for other 3-D geometries are covered. This historical account and literature survey of the crossflow instability are used for references prior to 1989. During the 0066-4189/03/0115-0413$14.00 413 Annu. Rev. Fluid Mech. 2003.35:413-440. Downloaded from arjournals.annualreviews.org by UNIVERSITY OF BATH on 05/19/10. For personal use only.

Upload: others

Post on 20-Jun-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM10.1146/annurev.fluid.35.101101.161045

Annu. Rev. Fluid Mech. 2003. 35:413–40doi: 10.1146/annurev.fluid.35.101101.161045

Copyright c© 2003 by Annual Reviews. All rights reserved

STABILITY AND TRANSITION OF

THREE-DIMENSIONAL BOUNDARY LAYERS

William S. Saric1, Helen L. Reed1, and Edward B. White21Mechanical and Aerospace Engineering, Arizona State University, Tempe, Arizona85287-6106; email: [email protected], [email protected] and Aerospace Engineering, Case Western Reserve University, Cleveland,Ohio 44106-7222; email: [email protected]

Key Words roughness, nonlinear, saturation, secondary instabilities, rotating disk,swept wing

■ Abstract The recent progress in three-dimensional boundary-layer stability andtransition is reviewed. The material focuses on the crossflow instability that leads totransition on swept wings and rotating disks. Following a brief overview of instabilitymechanisms and the crossflow problem, a summary of the important findings of the1990s is given.

1. INTRODUCTION

In low-disturbance environments such as flight, boundary-layer transition to tur-bulence generally occurs through the uninterrupted growth of linear instabilities.The initial conditions for these instabilities are introduced through the receptiv-ity process, which depends on a variety of factors (Saric et al. 2002). There isno shortage of publications in the field of boundary-layer stability and transi-tion. Comprehensive reviews for both two- and three-dimensional (3-D) flows aregiven by Arnal (1994), Mack (1984), and Reshotko (1994). Reed et al. (1996)give an up-to-date discussion of the effectiveness and limitations of linear the-ory in describing boundary-layer instabilities. The reader is referred to these re-ports for overviews of much of the early work in stability and transition. Crouchand coworkers (Crouch 1997, Crouch & Ng 2000, Crouch et al. 2001) andHerbert (1997b) have presented analyses of transition with specific applicationsto flight.

Reed & Saric (1989) review the early theoretical and experimental work in 3-Dboundary layers. Swept wings, rotating disks, axisymmetric bodies (rotating conesand spheres), corner flows, attachment-line instabilities, as well as the stability offlows for other 3-D geometries are covered. This historical account and literaturesurvey of the crossflow instability are used for references prior to 1989. During the

0066-4189/03/0115-0413$14.00 413

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 2: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

414 SARIC ¥ REED ¥ WHITE

past 12 years, most of the progress has occurred in rotating-disk and swept-wingflows, and this review concentrates on these areas.

The study of 3-D boundary layers is motivated by the need to understand thefundamental instability mechanisms that cause transition in swept-wing flows.Research has identified four types of instabilities for these flows: attachment line,streamwise, centrifugal, and crossflow. The attachment-line boundary layer can un-dergo an instability or be subject to contamination by wing-root turbulence; thesephenomena are associated with, in general, swept wings with a large leading-edgeradius (Hall et al. 1984, Reed & Saric 1989). The streamwise instability givesthe familiar Tollmien-Schlichting (T-S) wave in two-dimensional flows (Reshotko1976, Mack 1984, Reed et al. 1996). Centrifugal instabilities appear over con-cave regions on the surface and result in the development of G¨ortler vortices(Saric 1994). The crossflow instability occurs in regions of pressure gradient onswept surfaces or on rotating disks. In the inviscid region outside the boundarylayer, the combined influences of sweep and pressure gradient produce curvedstreamlines at the boundary-layer edge. Inside the boundary layer, the stream-wise velocity is reduced, but the pressure gradient is unchanged. Thus, the bal-ance between centripetal acceleration and pressure gradient does not exist. Thisimbalance results in a secondary flow in the boundary layer, called crossflow,that is perpendicular to the direction of the inviscid streamline. The 3-D pro-file and resolved streamwise and crossflow boundary-layer profiles are shown inFigure 1.

Because the crossflow velocity must vanish at the wall and at the edge of theboundary layer, an inflection point exists and provides a source of an inviscidinstability. The instability appears as co-rotating vortices whose axes are alignedto within a few degrees of the local inviscid streamlines.

Unlike T-S instabilities, the crossflow problem exhibits amplified disturbancesthat are stationary as well as traveling. Even though both types of waves arepresent in typical swept-wing or rotating-disk flows, transition is usually causedby either one, but not both, of these waves. Although linear theory predicts thatthe traveling disturbances have higher growth rates, transition in many exper-iments is induced by stationary waves. Whether stationary or traveling wavesdominate is related to the receptivity process. Stationary waves are more im-portant in low-turbulence environments characteristic of flight, whereas travelingwaves dominate in high-turbulence environments (Bippes 1997, Deyhle & Bippes1996).

Stationary crossflow waves (v′ andw′ disturbances) are typically very weak;hence, analytical models have been based on linear theory. However, experimentsoften show evidence of strong nonlinear effects (Dagenhart et al. 1989, 1990;Bippes & Nitschke-Kowsky 1990; Bippes et al. 1991; Deyhle et al. 1993; Reibertet al. 1996). Because the wave fronts are fixed with respect to the model and arenearly aligned with the potential-flow direction (i.e., the wavenumber vector isnearly perpendicular to the local inviscid streamline), the weak (v′, w′) motion ofthe wave convectsO(1) streamwise momentum producing a strongu′ distortion

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 3: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 415

Figure 1 Swept-wing boundary-layer profiles.

in the streamwise boundary-layer profile. This integrated effect and the resultinglocal distortion of the mean boundary layer lead to the modification of the basicstate and the early development of nonlinear effects.

An interesting feature of the stationary crossflow waves is the creation of sec-ondary instabilities. Theu′ distortions created by the stationary wave are timeindependent, resulting in a spanwise modulation of the mean streamwise velocityprofile. As the distortions grow, the boundary layer develops an alternating patternof accelerated, decelerated, and doubly inflected profiles. The inflected profilesare inviscidly unstable and, as such, are subject to a high-frequency secondaryinstability (Kohama et al. 1991, Malik et al. 1994, Wassermann & Kloker 2002).

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 4: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

416 SARIC ¥ REED ¥ WHITE

This secondary instability is highly amplified and leads to rapid local breakdown.Because transition develops locally, the transition front is nonuniform in span andcharacterized by a “saw-tooth” pattern of turbulent wedges.

Arnal (1997), Bippes (1999), and Kachanov (1996) review the European contri-butions to stability and transition in 3-D boundary layers and, as such, are compan-ion papers to this work. Recently, improvements in both experimental techniquesand computational methods have opened the door to a new understanding of transi-tion in 3-D boundary layers. This review focuses on the latest developments, withemphasis on the experimental work and relevant comparisons with computationalfluid dynamics (CFD).

2. ROTATING DISK

The boundary layer on the surface of a rotating disk has been used as a modelproblem for swept-wing flow because the velocity profiles on the disk provide abasic state and stability behavior very similar to that of the swept-wing boundarylayer shown in Figure 1. The advantage of the rotating disk is the existence of awell-known similarity solution for the basic flow that features a boundary layerof constant thickness. Reed & Saric (1989) reviewed in detail the progress on therotating-disk problem through the 1980s. At the time of that review, a number ofanalyses and experiments were beginning to provide a good picture of the linearstability of stationary waves, especially those connected to the upper inviscidbranch of the neutral stability curve, but much remained to be determined aboutthe lower viscous branch, traveling waves, and nonlinear behavior.

Linear stability of traveling waves was addressed by Balakumar & Malik (1991),who used a parallel-flow approximation and considered the neutral curves atnonzero frequencies. They found that, as the frequency increases, the Reynoldsnumber of the lower minimum decreases while that of the upper minimum in-creases. The minimum critical Reynolds number over all frequencies is found tobe a lower-branch minimum. For the frequencies examined, the Reynolds num-ber and wave angle associated with the upper-branch minimum do not changesignificantly, but there are large changes at the lower-branch minimum.

For the linear problem, Faller (1991) considered both the upper and lowerbranches of the rotating-disk neutral curve and found minimum critical Reynolds-number behavior equivalent to that found by Malik (1986) for the stationary upper-branch mode and similar to the results obtained by Balakumar & Malik (1991) forthe traveling lower-branch mode. Faller (1991) explained that the lower-branchmodes are typically not observed in experiments both because they have a lowergrowth rate than those on the upper branch and because, among the lower-branchmodes, stationary modes are much less amplified than traveling modes. This isin contrast to the upper branch where the amplification of stationary waves isnearly equal to the most amplified traveling waves. Because most experiments arebetter able to detect stationary waves than traveling waves, it is natural that slowly

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 5: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 417

growing traveling disturbances like the lower-branch modes can easily escapedetection.

Jarre et al. (1996a) measured traveling and stationary waves on a rotating diskto confirm linear-stability theory (LST) predictions, particularly those concerningtraveling waves. The fluctuation intensity measured close to the upper-branchstationary-wave neutral point is 2%, so the experiments have a relatively highdisturbance level. Jarre et al. (1996a) found a minimum critical Reynolds numberfor stationary waves reasonably close to the theoretical predictions. The agreementbetween the measurements and theory for the corresponding wavenumber and thedisturbance-mode eigenfunctions (Malik 1986) is not as good. This appears to bedue to the basic-state circumferential velocity not going to zero above the disk asthe similarity solution requires.

A somewhat different approach to the rotating-disk problem is taken in a se-ries of papers by Itoh (1994, 1996, 1997, 1998), Buck & Takagi (1997), Takagiet al. (1998, 2000), and Buck et al. (2000). Itoh (1994) formulated the linear-stability problem to explicitly include the curvature of the inviscid streamline atthe boundary-layer edge. He found that strongly curved inviscid streamlines pro-duce a centrifugal-type instability similar to the G¨ortler instability on concavewalls (Saric 1994). Because streamlines are highly curved near the symmetryaxis in rotating disk flow, Itoh (1996, 1998) investigated whether the streamline-curvature instability can occur in these boundary layers. He found that the linear-stability neutral curve has two lobes: one for the crossflow instability and theother for the streamline-curvature instability. The streamline-curvature instabil-ity is unstable at much lower Reynolds numbers than crossflow, and the rangeof unstable wave numbers decreases as the Reynolds number grows. For cross-flow, the unstable wave-number range increases with increasing Reynolds numberpast the minimum critical value. Itoh (1996, 1998) concluded that the streamline-curvature instability is the source of the lower branch of the neutral curve discussedabove.

Although the streamline-curvature instability is destabilized at a much lowerReynolds number than the crossflow instability, in most experiments the crossflowinstability is observed. This discrepancy is likely a result of either smaller growthrates or smaller receptivity of the streamline-curvature instability versus the cross-flow instability. To address the growth-rate question, Itoh (1997) calculated that,in a region unstable to both instabilities, the growth of streamline-curvature modesis larger than that of crossflow modes. In a subsequent experiment, Takagi et al.(1998) confirmed, at least qualitatively, the predictions by Itoh (1997). Experi-mental verification is also provided by Buck & Takagi (1997) and Takagi et al.(2000), who used point-source disturbance generators. Buck et al. (2000) alsodemonstrated that for point-source forcing early growth of streamline-curvaturedisturbances can suppress the appearance of crossflow disturbances that are ob-served under zero-forcing conditions. However, Buck et al. (2000) suggested thatthe suppression may not prevent or delay transition because the large-amplitudestreamline-curvature disturbances may lead to turbulence.

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 6: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

418 SARIC ¥ REED ¥ WHITE

2.1. Nonlinearities, Wave Interactions, andSecondary Instabilities

When the review by Reed & Saric (1989) was published there was only a limitedbody of work that considered the nonlinear behavior of finite-amplitude distur-bances in rotating-disk boundary layers. Subsequently, there have been severalanalytical and experimental investigations of the nonlinear regime. One of theanalytical works was by Vonderwell & Riahi (1993), who examined the weaklynonlinear behavior of waves near the minimum critical Reynolds number of theupper-branch neutral point. They developed a Landau-type equation for the waveamplitude and found that the Landau constant is negative such that there is a subcrit-ical instability atR<Rcand that disturbance amplitudes go to infinity (in the weaklynonlinear model) in finite time forR>Rc. This conflicts with the earlier finding byItoh (1985) stating that the instability is supercritical. Jarre et al. (1996b) conducteda follow-up experiment to their previous work (Jarre et al. 1996a) that determinedthe Landau constant from data. The experiment features a single, large-amplituderoughness element that generates finite-amplitude stationary waves. Even thoughthe basic state differs from that of the similarity solution as mentioned above, theresults do seem to support at least a qualitative analysis of the Landau equationcoefficients. In contrast to the results presented by Vonderwell & Riahi (1993) thatindicated a negative Landau constant, Jarre et al. (1996b) found that the Landauconstant is positive, meaning that the instability is supercritical. This result seemsreasonable in light of the related swept-wing experiments discussed below thatshow that finite-amplitude crossflow waves undergo amplitude saturation.

Corke & Knasiak (1998) considered nonlinear wave interactions on a rotatingdisk. They used a polished disk with various periodic roughness configurationslocated near the minimum critical Reynolds number,Rc, for upper-branch modes.Various tests showed that stationary modes obeyed linear theory in terms of circum-ferential wave number, growth rate, and mode shape up to a radius ofR/Rc≈ 1.5.Traveling waves are also well predicted by linear theory. In the nonlinear regime,Corke & Knasiak (1998) found significant growth of very-low-wave-number sta-tionary waves that are obviously not growing according to linear theory. Using thebicoherence of traveling and stationary wave phases, they confirmed a triad reso-nance. In all of the roughness configurations tested (including no artificial rough-ness), low circumferential wave numbers were produced by triad resonance andgrew to at least as large an amplitude as the most amplified stationary disturbances.

A mechanism for the final stage of the transition process, breakdown of thelaminar boundary layer by means of a secondary instability, was considered byBalachandar et al. (1992) using a Floquet approach that considered a steady basicstate in the rotating reference frame consisting of the mean flow plus the mostamplified stationary disturbance predicted by linear theory at amplitudes rangingfrom 6% to 12%. Below 9% amplitude, no secondary modes are destabilized, butabove this threshold, a broad band of secondary modes are destabilized with growthrates as high as four times that of the stationary waves of the primary instability.

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 7: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 419

2.2. Absolute Instabilities

Perhaps the most important recent development in the study of the rotating-diskproblem is the theory by Lingwood (1995) that rotating-disk boundary layerssupport an absolute instability atR = 510. The implication of this finding is thatthe steady, laminar basic state cannot exist beyond this Reynolds number regardlessof the care taken in preparing and conducting the experiment. This finding is ingood agreement with various rotating-disk experiments, most of which do notmaintain laminar flow much beyondR = 500.

Lingwood (1995) followed Briggs’ method (see Huerre & Monkewitz 1990),which prescribes the means to evaluate the Fourier-Laplace integral that arisesfrom considering an initial-boundary-value problem for impulsive forcing in adeveloping flow. If the group velocity of a disturbance wave packet goes to zerowhile the temporal growth rate is positive, then the flow is absolutely unstable.Lingwood (1995) applied Briggs’ method by using a parallel-flow approximationand observed a pinch between two neutral branches of the dispersion relationshipthat yields a positive temporal growth rate at a Reynolds numberR≥ 510.625.As R is increased, the absolutely unstable wave-number range expands and atlarge R tends toward the limits found for the inviscid case. Lingwood (1995)concluded that the absolute instability arises from an inviscid mechanism and thatthe good correlation between the Reynolds-number limit for absolute instabilityand the experimental transition Reynolds numbers makes a case for transitionbeing triggered by the absolute instability.

Lingwood (1996) then performed an experiment on a rotating disk in whichstationary-wave development leading to transition was observed with both an un-excited and an excited flow in which a once-per-revolution transient disturbancewas introduced near the center of the disk. The unexcited flow exhibited stationarywaves that grew in a manner consistent with LST and led to transitional behaviorby R = 514 with fully turbulent behavior observed by approximatelyR = 600.Perhaps the most important information from the tests without excitation is thatthe spectra of the time series do not show any evidence of secondary instabilitiesprior to breakdown. Balachandar et al. (1992) claimed that an amplitude of at least9% would be required to see such behavior and that Lingwood’s (1996) stationarywaves only reach 3% prior to transition, so this result is not unexpected. The otherinteresting feature of the spectra is that there is an abrupt jump of approximately afactor of 10 in the fluctuation amplitudes at all frequencies whenR is between 497and 514 that is not preceded by any significant signs of imminent breakdown to tur-bulence. In fact, downstream of the jump, the spectra do not appear turbulent; thereis a distinct spectral peak associated with the stationary waves and a lower am-plitude peak associated with harmonics of the stationary waves. The jump occursexactly in the region of the absolute instability identified by Lingwood (1995).

A more compelling demonstration of the existence of the absolute instabilitycomes from Lingwood’s (1996) experiments that include transient forcing. Theleading edge of the wave packet convects in the radial direction at an essentially

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 8: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

420 SARIC ¥ REED ¥ WHITE

constant rate, whereas the propagation rate of the trailing edge decreases dra-matically as the packet approaches the critical Reynolds number for the absoluteinstability, as predicted by Lingwood’s earlier analysis. Taken together, the ex-perimental evidence convincingly shows that for cases with low-amplitude initialdisturbances (i.e., low enough to avoid secondary instabilities) disturbances passfrom a convective-instability regime forR< 510 to an absolutely unstable regimefor R> 510. The discovery that an absolute instability exists on rotating disks isimportant because this places a firm upper bound on the Reynolds numbers forwhich laminar flow is possible.

The improved picture of rotating-disk instabilities includes a number of routesto turbulence. First, high-amplitude traveling disturbances (if they exist) may beamplified at very low Reynolds numbers. At higher Reynolds numbers, stationarywaves dominate. These can lead to turbulence through two distinct mechanisms.If the disturbance levels are high, then high-frequency secondary instabilities leadto rapid transition. If disturbance levels are low and secondary instabilities donot appear, the flow becomes absolutely unstable for Reynolds numbers above510, and transition occurs. Despite a better understanding of this basic scenario,much remains to be learned, especially in regard to nonlinear behavior. Oneimportant problem that remains is to conclusively identify whether the variousbranches of the instability are subcritical or supercritical. Another unresolvedquestion is what effect, if any, finite-amplitude disturbances have on the absoluteinstability.

3. SWEPT WING

3.1. Swept Flat Plate

A prototype of the swept wing is a swept flat plate with a pressure body (Saric &Yeates 1985) that could be used to verify the basic aspects of LST. The advantageof the flat plate is that the first neutral point can be placed away from the leadingedge and thus avoiding nonparallel and curvature effects. In a series of experimentson swept flat plates (Nitschke-Kowsky & Bippes 1988, M¨uller 1990, Deyhle et al.1993, Kachanov & Tararykin 1990, Kachanov 1996), the observed wavelengthand growth rate of the crossflow wave is in general agreement with linear theory(Orr-Sommerfeld solutions).

Deyhle et al. (1993) and Kachanov (1996) developed techniques to create con-trolled traveling waves within the boundary layer and observed the growth ofthem. Linear theory was also verified in this case. Bippes (1997) put into per-spective the importance of traveling modes and their dependence on freestreamconditions. In general, under conditions of high freestream turbulence, travelingmodes dominate and linear theory predicts the behavior of the crossflow waves. Inlow-turbulence flows, stationary waves dominate, the mean flow is distorted, andnonlinear behavior occurs rapidly, thus obviating the linear analysis beyond initialmode selection.

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 9: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 421

3.2. Receptivity

When the paper by Reed & Saric (1989) was published, very little was knownregarding receptivity of 3-D boundary layers. It was acknowledged that recep-tivity must play a major role in determining the details of crossflow transitionbecause of the wide range of behaviors observed using different models in variousexperimental facilities. In detailing a list of unanswered questions regarding thesevariations observed in different facilities, Reed & Saric (1989) concluded, “All ofthis serves notice that stability and transition phenomena are extremely dependenton initial conditions.” Since then, a significant research effort that includes exper-imental, theoretical, and computational work has made great progress toward anunderstanding of the applicable receptivity processes [the reader is referred to thereview by Bippes (1999) for additional details]. Owing to space constraints, wefocus primarily on the recent efforts in this area.

ROLE OF FREESTREAM FLUCTUATIONS The effect of freestream turbulence oncrossflow transition was investigated by Deyhle & Bippes (1996), who performedtransition measurements on a crossflow-dominated swept-plate model in a numberof different wind-tunnel facilities with varying freestream-turbulence levels. Theyfound that for the particular model used in those experiments turbulence intensitiesaboveTu = 0.0015 produced transition behavior dominated by traveling waves,but that for lower turbulence levels stationary waves dominated. It is surprising tonote that for increased turbulence levels where traveling waves dominate but theturbulence intensity is not too high, 0.0015< Tu< 0.0020, transition was actu-ally delayed relative to low-turbulence cases at the same Reynolds number. Theexplanation for this is that the traveling waves excited by the increased freestreamturbulence were sufficiently strong to prevent stationary waves from causing tran-sition but were not strong enough to cause transition as quickly as the stationarywaves they replaced. This behavior indicates that transition results from manywind tunnels may have no bearing on flight results because quite low levels ofturbulence are sufficient to generate traveling-wave-dominated behavior, counterto the stationary-wave-dominated behavior observed in flight.

In another experiment, Radeztsky et al. (1999) found that transition behavioron a swept wing is insensitive to sound, even at amplitudes greater than 100 dB.The conclusion is that the variations observed by Deyhle & Bippes (1996) atvarying levels ofTu are due primarily to variations in the vortical components ofthe freestream fluctuations and not to the acoustic component.

ROLE OF SURFACE ROUGHNESS The receptivity mechanism for the stationary vor-tices that are important for transition in environments with very-low-amplitudeturbulent fluctuations (i.e., characteristic of the flight environment) is surfaceroughness. This was conclusively established by M¨uller & Bippes (1989), whotranslated a swept-flat-plate model relative to a wind-tunnel test section and foundthat the recurring stationary transition pattern was fixed to the model. The in-stability features they observed had to be related to model roughness rather than

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 10: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

422 SARIC ¥ REED ¥ WHITE

to fixed features of the freestream flow generated by nonuniformities of the screensor other effects.

Detailed roughness studies of isolated 3-D roughness features have been com-pleted by Juillen & Arnal (1990), who found that for isolated roughness elementsthe correlation by von Doenhoff & Braslow (1961) describing the limit for bypasstransition is correct. Radeztsky et al. (1999) showed that the characteristics ofisolated 3-D roughness elements whoseRek values fall below the bypass-inducinglevel play a very important role in transition behavior. Radeztsky et al. (1999)found that roughness is most effective at generating crossflow disturbances at orjust upstream of the first neutral point, that the transition location is quite sensitiveto roughness height even for roughness Reynolds numbers as low asRek = 0.1,and that the roughness diameter must be greater than 10% of the most amplifiedstationary wavelength to be effective.

In addition to isolated 3-D roughness, natural surface roughness can also play asignificant role in transition location. Radeztsky et al. (1999) found that a decreasein surface-roughness amplitude from 9.0µm rms to 0.25µm rms increases thetransition Reynolds number by 70%. In contrast to this, experiments by Reibertet al. (1996) showed that an artificial distributed roughness array with an amplitudeof 6 µm or greater applied near the leading edge produces transition behavioralmost completely insensitive to roughness amplitude in the range of 6–50µm. Inthe experiment by Reibert et al. (1996), the periodic roughness arrays concentratethe disturbances in a very narrow band of wavelengths, and these well-definedmodes generally lead to nonlinear amplitude saturation of the most amplifiedwave. Amplitude saturation renders the initial amplitude nearly unimportant solong as it is above a threshold for which saturation occurs prior to transition.In other words, Radeztsky et al. (1999) saw a strong roughness-amplitude effectfor roughness conditions that did not lead to saturation, whereas Reibert et al.(1996) did not see a strong roughness-amplitude effect for roughness at or abovesaturation-producing amplitudes.

RECEPTIVITY COMPUTATIONS A number of theoretical and computational ap-proaches to swept-wing crossflow receptivity have been applied. Some of themore recent include an adjoint equation approach by Fedorov (1989), a parabo-lized stability equation (PSE; linear PSE, LPSE; nonlinear PSE, NPSE) approachby Herbert & Lin (1993), and a direct numerical simulation (DNS) approach bySpalart (1993). Other efforts are by Crouch (1993, 1994, 1995) and Choudhari(1994), both consider the receptivity of Falkner-Skan-Cooke boundary layers asperturbations of a parallel boundary layer. The framework of their approaches al-lows both surface roughness and freestream acoustic disturbances to be consideredas disturbance sources. Choudhari (1994) extended his work to include acoustic-wave-angle effects and a variety of different roughness configurations includingisolated roughness, roughness arrays and lattices, as well as distributed randomroughness. Crouch (1994) emphasized a framework equally applicable to T-S andcrossflow disturbances. Because traveling-wave receptivity scales with two small

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 11: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 423

parameters (the freestream velocity-fluctuation amplitude and surface-roughnessamplitude), whereas the stationary-wave receptivity scales with only one (the sur-face roughness), both authors note that stationary waves should be expected todominate in low-disturbance environments and that traveling waves should onlyappear for large freestream fluctuation levels. The experiments by Radeztsky et al.(1999) confirm that acoustic forcing is not an effective receptivity source when thesurface roughness is low.

The method described by Crouch (1994) is used by Ng & Crouch (1999) tomodel the artificial roughness arrays used in the Arizona State University (ASU)swept-wing experiments by Reibert et al. (1996). Ng & Crouch (1999) give resultsthat are in good agreement with the experiments when the receptivity of 12-mmwaves (most unstable mode) to low-amplitude (roughness height 6 or 18µm),12-mm-spaced roughness arrays is considered. However, the linear theory over-predicts the receptivity of the 48-µm roughness, suggesting the nonlinear effectsreduce receptivity effectiveness. Results are not as good for the receptivity of9-, 12-, and 18-mm waves to 36-mm-spaced roughness, but this problem maybe complicated by the nonlinear growth and production of harmonics associatedwith the stationary crossflow waves. Bertolotti (2000) presented another Fourier-transform approach that includes the effects of basic flow nonparallelism. Bertolottiapplied the method to both swept Hiemenz flow and the Deutschen Centrum f¨urLuft- und Raumfhahrt (DLR) swept-flat-plate experiment by Bippes and cowork-ers. Comparisons between the receptivity predictions and the DLR swept-plateresults are quite good.

Another recent approach by Collis & Lele (1999) consists of solving the steadyNavier-Stokes equations in the leading-edge region of a swept parabolic bodyand of using that solution as a basic state for a linearized steady-disturbance sys-tem that includes surface roughness. Comparing the results of this approach tothose obtained by Choudhari (1994) and Crouch (1994) shows that receptivity tosurface roughness is enhanced by convex surface curvature and suppressed by non-parallelism. Neglecting nonparallelism causes the local approach to overpredictreceptivity by as much as 77% for the most amplified stationary crossflow wave.The error introduced by neglecting nonparallelism is most severe for wavelengthsin the range most amplified by the crossflow instability and for roughness closeto the first neutral point. Janke (2001) confirmed overprediction of receptivity bylocal approaches. The implication is that amplitude-based transition-predictionmethods need to employ a receptivity model that includes nonparallelism becausethe crossflow modes that dominate transition are most strongly affected by this in-fluence. This result stands somewhat in contradiction to the result by Ng & Crouch(1999) that includes neither surface curvature nor nonparallelism yet shows goodagreement with experimental results.

ROLE OF TURBULENCE/ROUGHNESS INTERACTIONS Although receptivity modelshave considered the interaction of surface roughness with acoustic fluctuations,none consider the receptivity of freestream turbulence interacting with surface

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 12: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

30 Nov 2002 9:49 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

424 SARIC ¥ REED ¥ WHITE

roughness. The experiments by Radeztsky et al. (1999) and Deyhle & Bippes(1996) would suggest however that the turbulent fluctuations play a much moresignificant role in the transition process than acoustic fluctuations. Deyhle &Bippes (1996) give a turbulence intensity criterion that selects traveling or station-ary modes, but drawing an analogy to the acoustic/roughness interaction results,one should suspect that it is the interaction of freestream turbulence with surfaceroughness that is the important consideration. Turbulence level alone may not besufficient to predict whether traveling or stationary waves will be most important.

With this in mind, the experiment by White et al. (2001) examined the in-teractions of surface roughness and freestream turbulence and confirmed that theselection of traveling-wave or stationary-wave behavior is not as straightforward astheTu = 0.0015 criterion. In that experiment, a swept-wing model outfitted witha variable-amplitude roughness insert atx/c = 0.025 (near the first crossflow neu-tral point) was used to change the roughness configuration during an experimentfrom nominally smooth (<0.5µm rms) to an 8-mm-spaced, 50-µm-high rough-ness pattern with 2-mm diameter roughness elements. A turbulence-generatinggrid provided a turbulence intensity ofTu = 0.003. With a nominally smoothleading edge, a sharp saw-tooth transition pattern was observed in a naphthaleneflow-visualization experiment, indicating stationary-wave-dominated transition,but when the roughness array was activated, the saw-tooth transition front wasreplaced by a diffuse spanwise-invariant transition front indicative of traveling-wave-dominated transition. The implication is that traveling waves do result froman interaction of freestream velocity fluctuations with surface roughness and notfrom turbulence intensity alone. Therefore, even though stationary waves are gener-ated by surface roughness, increasing the roughness amplitude does not necessarilymake stationary waves more likely to be observed than traveling waves. Instead,both surface roughness and turbulence intensity must be considered together in amore sophisticated manner.

3.3. Nonlinear Behavior

NONLINEAR SATURATION In all the experiments by Bippes and coworkers, thegrowth of the stationary and traveling crossflow waves showed initial qualitativeagreement with linear theory. However, the disturbance amplitude saturated owingto nonlinear effects. Also, the amplitude of the traveling waves showed a span-wise modulation indicating nonlinear interactions with the stationary modes. Thestreamwise-mode shapes had saturation amplitudes of approximately 20%.

Reibert et al. (1996) investigated the nonlinear saturation of stationary wavesusing micron-sized artificial roughness elements to control the initial conditions.Full-span arrays of roughness elements were used to preserve the uniform spanwiseperiodicity of the disturbance as shown in the velocity contour plot of Figure 2,where the crossflow moves from left to right.

The development of crossflow occurs in two stages. The first stage is linearand is characterized by small verticalv′ and spanwisew′ disturbance velocities

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 13: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 425

convecting low-momentum fluid away from the wall and high-momentum fluidtoward the wall. This exchange of momentum occurs in a region very close to thewall where there are large vertical gradients in the basic-state streamwise velocity.Because of this large gradient, the small displacements caused by thev′ andw′

disturbance components quickly lead to large disturbancesu′ superposed on thebasic state further downstream. Thisu′ component soon becomes too large, andnonlinear interactions must be included in any calculations. This is the secondstage, evidenced by rollover seen in the streamwise-velocity contours. By forc-ing the most unstable mode (according to linear theory), nonlinear saturation ofthe disturbance amplitude is observed well before transition. Although the initialgrowth rate increases with increasing roughness height, the saturation amplituderemains largely unaffected by changes in the roughness height. The presence ofa large laminar extent of nonlinear saturation gives rise to a certain difficulty inusing linear methods (such aseN or LPSE) to predict transition. The futility ofsuch approaches is expressed by Arnal (1994) and Reed et al. (1996), who showthat linear methods do not work in correlating transition for crossflow-dominatedboundary layers.

MODAL DECOMPOSITION Radeztsky et al. (1994) described a measurement tech-nique that allows the experimentally obtained stationary crossflow structure to bedecomposed into its spatial modes. Using a high-resolution traversing mechanism,hotwires are carefully moved through the boundary layer along a predeterminedpath. Data are acquired at numerous spanwise locations, from which modal in-formation is extracted using a spatial power spectrum. Reibert et al. (1996) useda slightly modified technique to more objectively determine the modal content.Under certain conditions, the amplitude of the fundamental disturbance mode andeight harmonics are successfully extracted from the experimental data.

EXCITATION OF LESS UNSTABLE MODES Using the modal decomposition tech-nique described above, Reibert et al. (1996) investigated the effect of roughness-induced forcing at a wavelength three times that of the most unstable stationarymode (according to linear theory). A cascading of energy from the fundamental tohigher modes (smaller wavelengths) was observed, leading to nonlinear interac-tions among the fundamental mode and its harmonics. Transition was observed tooccur slightly earlier compared to forcing at the most unstable wavelength, and thesaw-tooth transition front was much more “regular” in span. These data indicatethat nonlinear interactions among multiple modes are important in determiningthe details of transition.

EXCITATION OF SUBCRITICAL MODES Continuing the experiments by Reibert et al.(1996), Saric et al. (1998) described a set of experiments in which the stationarycrossflow disturbance is forced with subcritical roughness spacing, i.e., the spac-ing between roughness elements is less than the wavelength of the most unstablemode. Under these conditions, the rapid growth of the forced mode completelysuppresses the linearly most unstable mode, thereby delaying transition beyond its

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 14: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

426 SARIC ¥ REED ¥ WHITE

“natural” location (i.e., where transition occurs in the absence of artificial rough-ness). These data demonstrate that surface roughness can be used to control thestationary crossflow disturbance wave-number spectrum in order to delay transi-tion on swept wings.

STRUCTURE IDENTIFICATION USING POD Chapman et al. (1998) applied linearstochastic estimation and proper orthogonal decomposition (POD) to identify thespatio-temporal evolution of structures within a swept-wing boundary layer. De-tailed measurements are acquired using multielement hotfilm, hotwire, and cross-wire anemometry. These data allow the POD to objectively determine (based onenergy) the modes characteristic of the measured flow. Data are acquired throughthe transition region, from which an objective transition-detection method isdeveloped using the streamwise spatial POD solutions.

CFD COMPARISONS (DNS) DNS have historically been constrained by computerresources and algorithmic limitations; however, some successes have been achievedin relation to the stationary crossflow problem. Reed & Lin (1987) and Lin & Reed(1993) performed DNS for stationary waves on an infinite-span swept wing simi-lar to the ASU experiments. Meyer & Kleiser (1990) investigated the disturbanceinteractions between stationary and traveling crossflow modes on a swept flat plateusing Falkner-Scan-Cooke similarity profiles for the basic state. The results arecompared to the experiments by M¨uller & Bippes (1989). With an appropriateinitial disturbance field, the nonlinear development of stationary and travelingcrossflow modes is simulated reasonably well up to transition. Wintergerste &Kleiser (1995) continued this work by using DNS to investigate the breakdown ofcrossflow vortices in the highly nonlinear final stages of transition.

CFD COMPARISONS (PSE) Combining the ability to include nonparallel and non-linear effects with computationally efficient parabolic marching algorithms, thePSE developed by Herbert (1997a) have recently been used to successfully modelthe crossflow instability. For swept-wing flows, NPSE calculations exhibit the dis-turbance amplitude saturation characteristic of the DLR and ASU experiments.Wang et al. (1994) investigated both stationary and traveling crossflow waves forthe swept airfoil used in the ASU experiments and predicted nonlinear amplitudesaturation for both types of disturbances. It is suggested that the interaction betweenthe stationary and traveling waves is an important aspect of the transition process.

Haynes & Reed (2000) recently validated the NPSE approach for 3-D flowssubjected to crossflow disturbances. Here a detailed comparison of NPSE resultswith the experimental measurements by Reibert et al. (1996) show remarkablygood agreement. Haynes & Reed (2000) independently computed the inviscidflow for the model, from which the edge boundary conditions were generated forthe boundary-layer code. The initial conditions for the NPSE calculation (withcurvature) were obtained by solving the local LST equations at 5% chord locationfor the fundamental stationary mode and adjusting its RMS amplitude such thatthe total disturbance amplitude matched that of the experiment at 10% chord. The

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 15: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

30 Nov 2002 9:51 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 427

Figure 3 Disturbance amplitude for Orr-Sommerfeld, LPSE, NPSE (Haynes & Reed2000), and experiments (Reibert et al. 1996).

NPSE was then matched from 5% chord to 45% chord. Transition occurred onthe experimental model at 47% chord. The primary and higher modes all growrapidly at first and saturate at approximately 30% chord. This is due to a strongnonlinear interaction among all the modes over a large chordwise distance. Aninteresting discovery by Haynes & Reed (2000) is that the flow is hypersensitive toseemingly negligible streamline and body curvature and that inclusion of curvatureterms is required to adequately match the experimental data. Figure 3 shows thecomparison of the experimentalN-factor (log of the amplitude ratio) with LPSE,NPSE, and LST as the crossflow vortex grows and then saturates downstream.All the computations include curvature. It is clear that the linear theories fail toaccurately describe saturation for this situation.

Figure 4 shows a comparison of the experimental and computational totalstreamwise velocity contours at 45% chord; the agreement between the NPSEand the experiments is excellent. The computations in Figures 3 and 4 use only aninitial amplitude. The basic state is computed independently from the experimentaldata.

There has been much debate about the effects of curvature. For the ASU config-uration, the inclusion of curvature has a very small effect on the metric coefficients.The maximum values of metric coefficients occur at 5% chord where they are of

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 16: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

428 SARIC ¥ REED ¥ WHITE

the order of 1× 10−2 and 10−3, respectively. They both drop off sharply withincreasing chordwise distance. These values may compel the researcher to ne-glect curvature, but the work by Haynes & Reed (2000) demonstrates conclusivelythat small changes in the metric coefficients can have a significant effect on thedevelopment of crossflow vortices.

Radeztsky et al. (1994) studied the effects of angle-of-attack. Here, in a case ofweak favorable pressure gradient, the experiments showed that the crossflow distur-bance is decaying in disagreement with various linear theories (LST, LPSE/withoutcurvature, and LST/with curvature) that predicted a growing disturbance.Radeztsky et al. (1994) concluded that the disagreement was due to nonlinear-ity. For this case, Haynes & Reed (2000) found that the LPSE/with curvature andNPSE/with curvature both agreed with the experiment, indicating that, in fact, thecrossflow disturbance decays and that there is a strong sensitivity to changes incurvature, nonparallel effects, and pressure gradient (angle-of-attack). The distur-bance was linear for this case.

3.4. Control with Distributed Roughness

Two important observations concerning the distributed roughness results byReibert et al. (1996) are (a) unstable waves occur only at integer multiples ofthe primary disturbance wave number and (b) no subharmonic disturbances aredestabilized. Spacing the roughness elements with wave numberk= 2π/λ apartexcites harmonic disturbances with spanwise wave numbers of 2k, 3k, . . . , nk (cor-responding toλ/2, λ/3, . . . , λ/n) but does not produce any unstable waves with“intermediate” wavelengths or with wavelengths greater thanλ.

Following this lead, Saric et al. (1998) investigated the effects of distributedroughness whose primary disturbance wave number does not contain a harmonicat λs = 12 mm (the most unstable wavelength according to linear theory). Bychanging the fundamental disturbance wavelength (i.e., the roughness spacing)to 18 mm, the velocity contours clearly showed the presence of the 18-, 9-, and6-mm wavelengths. However, the linearly most unstable disturbance (λs= 12 mm)has been completely suppressed. Moreover (and consistent with all previousresults), no subharmonic disturbances are observed, which shows that an appropri-ately designed roughness configuration can, in fact, inhibit the growth of the (natu-rally occurring) most unstable disturbance. When the disturbance wavelength wasforced at 8 mm, the growth of all disturbances of greater wavelength was suppres-sed. The most remarkable result obtained from the subcritical roughness spacingis the dramatic effect on transition location: In the absence of artificial rough-ness, transition occurs at 71% chord. Adding roughness with a spanwise spac-ing equal to the wavelength of the linearly most unstable wave moves transitionforward to 47% chord. However, subcritical forcing at 8-mm spanwise spacing ac-tually delays transition beyond the pressure minimum and well beyond 80% chord(the actual location was beyond view). This promising technique is currently beingevaluated for supersonic flight (Saric & Reed 2002).

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 17: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 429

Subsequent to the experiments, the NPSE results (Haynes & Reed 2000) con-firmed this effect. In a DNS solution, Wassermann & Kloker (2002) have shown thesame stabilization due to subcritical forcing. Using the same independent approachin terms of the calculation of the basic state, they demonstrated the stabilizationdue to subcritical roughness and coined the name transition delay by “upstreamflow deformation.”

3.5. Secondary Instabilities

Once stationary vortices reach saturation amplitude, this state can persist for asignificant streamwise distance. The velocity contours of Figures 2 and 4 showlow-momentum fluid above high-momentum fluid, which produces a double in-flection point in the wall-normal velocity profile. There is also an inflection pointin the spanwise profile. These inflection points are high in the boundary layer, andthe saturated vortices become unstable to a high-frequency secondary instabilitythat ultimately brings about transition to turbulence. Because of the importanceof the secondary instabilities in determining the location of breakdown of thelaminar flow, there have been a number of investigations, both experimental andcomputational, in this area. Bippes (1999) included details on the German efforts,in particular, the work by Lerche (1996) that emphasizes secondary instabilitiesin flows with higher turbulence levels and traveling crossflow waves. Boiko et al.(1995, 1999) covered recent efforts involving secondary instabilities in the Russiantraveling-wave experiments.

Poll (1985) conducted the first crossflow experiment for which a high-frequencydisturbance was observed prior to transition. Traveling crossflow waves were ob-served with a dominant frequency of 1.1 kHz forRec = 0.9× 106. Increasing thechord Reynolds number to 1.2× 106 increased the traveling crossflow frequencyto 1.5 kHz and also included an intermittent signal at 17.5 kHz superposed onthe underlying traveling crossflow waves. Poll (1985) noted that increasing theReynolds number beyond 1.2× 106 resulted in turbulent flow at the measurementlocation, so the high-frequency signal appeared only in a narrow range just priorto transition. Poll (1985) attributed the existence of the high-frequency componentto intermittent turbulence.

Kohama et al. (1991) investigated a high-frequency secondary instability specif-ically as a source of breakdown. This experiment combined hotwire measurementsand flow visualizations and was intended to determine the location and behaviorof the secondary instability mode relative to visualized breakdown patterns. TheKohama et al. (1991) experiments clearly show that there is a growing high-frequency mode in the region upstream of transition that can be associated with aninviscid instability of the distorted mean flow. However, a concern can be raised be-cause the measurements were made without a well-controlled primary disturbancestate. Experiments subsequent to this work used arrays of micron-sized roughnesselements near the leading edge that established the spanwise uniformity both ofthe stationary vortex amplitudes and of the transition location. Without the benefit

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 18: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

430 SARIC ¥ REED ¥ WHITE

of this technique, the data obtained by Kohama et al. (1991) likely spanned a widerange of stability behavior despite having been obtained at a single-chord posi-tion. Improvements in experimental techniques mean that more recent secondaryinstability experiments have replaced the work by Kohama et al. (1991) as the bestsource for secondary instability data.

Kohama et al. (1996) provided somewhat more detail than Kohama et al. (1991)by including velocity fluctuation maps that are filtered to give either primary insta-bility or secondary instability fluctuation levels. Kohama et al. (1996) concludedthat a “turbulent wedge starts from the middle height of the boundary layer” andthat this behavior is different from the usual picture of a turbulent wedge that orig-inates in the high-shear regions in naphthalene flow-visualization experiments.A subsequent swept-plate experiment by Kawakami et al. (1999) was conductedto further refine these measurements. This experiment featured a small speakermounted flush with the surface that permitted tracking of particular secondary in-stability frequencies. Without acoustic forcing, two separate high-frequency bandsof disturbances were observed to be unstable. At a chord Reynolds number of4.9 × 106, a band located between 600 and 2.5 kHz destabilized just downstreamof x/c = 0.35, and a second band located between 2.5 and 4.0 kHz destabi-lized just upstream ofx/c = 0.50. Transition was observed aroundx/c = 0.70.With acoustic forcing applied, the secondary instability frequency with the largestgrowth betweenx/c = 0.40 andx/c = 0.475 was observed to be 1.5 kHz.

In an effort to provide a more concrete experimental database on the behaviorof the secondary instability, White & Saric (2003) conducted a very detailed ex-periment that tracked the development of secondary instabilities on a swept wingthroughout their development for various Reynolds numbers and roughness con-figurations. They found a number of unique secondary instability modes that canoccur at different frequency bands and at different locations within the stationaryvortex structure. In this experiment, as many as six distinct modes are observedbetween 2 and 20 kHz. The lowest-frequency mode is nearly always the highestamplitude of all the secondary instabilities and is always associated with an ex-tremum in the spanwise gradient,∂U/∂z, which Malik et al. (1996, 1999) referto as a mode-I orzmode. Higher-frequency modes include both harmonics of thelowest-frequencyz mode that appear at the same location within the vortex anddistinct mode-II ory modes that form in the∂U/∂y shear layer in the portion ofthe vortex farthest from the wall. The lowest-frequency mode is typically detectedupstream of any of the higher-frequency modes. However, many higher-frequencymodes appear within a very short distance downstream. All of the secondary in-stability modes are amplified at a much greater rate than the primary stationaryvortices (even prior to their saturation). The rapid growth leads very quickly to thebreakdown of laminar flow, within∼5% chord of where the secondary instabilityis first detected. A consequence of this for transition-prediction methodologiesis that adequate engineering predictions of transition location could be obtainedfrom simply identifying where the secondary instabilities are destabilized becausethey lead to turbulence in such a short distance downstream of their destabilization

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 19: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 431

location. An interesting feature of the breakdown of the stationary vortex struc-ture is that it is highly localized. Spectra obtained by White & Saric (2003) atvarious points within the structure indicate that the first point to feature a broad,flat velocity-fluctuation spectrum characteristic of turbulence is very close to thewall in the region of highest wall shear. Other points in the structure remain es-sentially laminar for some distance downstream of the initial breakdown location.This finding supports the notion of a turbulent wedge originating near the wall, incontrast to what Kohama et al. (1996) concluded.

A successful computational approach to the secondary instability was presentedby Malik et al. (1994), who used an NPSE code to calculate the primary instabilitybehavior of stationary disturbances of a swept Hiemenz flow. As described pre-viously, the NPSE approach successfully captures the nonlinear effects includingamplitude saturation. The distorted meanflow provides a basic state for a local,temporal secondary instability calculation. The most unstable frequency is approx-imately one order of magnitude greater than the most unstable primary travelingwave, similar to the results by Kohama et al. (1991), and the peak mode amplitudeis “on top” of the stationary crossflow vortex structure. This location correspondsto what Malik et al. (1996) (see below) referred to as the mode-II secondaryinstability.

In order to obtain a more direct comparison to experimental data, Malik et al.(1996) used parameters designed to match the conditions found for the swept-cylinder experiment by Poll (1985) and the swept-wing experiment by Kohamaet al. (1991). The calculations by Malik et al. (1996) revealed that the energyproduction for a mode-I instability is dominated by the term〈u2w2〉∂U2/∂z2 andthe mode-II instability is dominated by〈u2v2〉∂U2/∂y2, where the subscript 2refers to a primary-vortex-oriented coordinate system. This energy-productionbehavior suggests that the mode-I instability is generated primarily by inflectionpoints in the spanwise direction and that the mode-II instability is generated byinflection points in the wall-normal direction. This situation is analogous to thesecondary instabilities of G¨ortler vortices (Saric 1994). Malik et al. (1996) claimedthat the fluctuations observed by Kohama et al. (1991) are mode-II instabilities, butthe spectral data presented by Kohama et al. (1991) likely includes contributionsof both the type-I and type-II modes. Although one or the other production mech-anism may dominate for a particular mode, it is too simplistic to assume that onlythe spanwise or wall-normal inflection points are responsible for the appearanceof a particular mode; with such a highly distorted 3-D boundary layer, all possibleinstabilities must be evaluated.

Malik et al. (1996) also computed the secondary instability behavior observedby Poll (1985) and predicted a 17.2-kHz mode compared to Poll’s high-frequencysignal, which occurred at 17.5 kHz. Based on the shape of this disturbance, Maliket al. (1996) claimed that this is a type-II mode. Malik et al. (1999) applied the sameapproach to the swept-wing experiments by Reibert et al. (1996). Malik et al. (1999)again applied a local, temporal stability of the stationary crossflow vortices that areestablished by the primary instability and found that better transition correlation

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 20: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

432 SARIC ¥ REED ¥ WHITE

results can be obtained by following the growth of the secondary instability in anN-factor calculation than by simply basing a prediction on the location at whichthe secondary instability destabilizes. A method based on the primary instabilityalone cannot adequately predict transition location.

Malik et al. (1999) found that a type-II mode becomes unstable uptream of anytype-I mode. This is not what was observed in the experiment by White & Saric(2003) in which type-I modes always appeared upstream of type-II modes. Thisapparent difference is likely due to the fact that freestream disturbance levels inthe experiment are stronger at frequencies near 3 kHz that correspond to the type-Imodes than at the higher frequencies near 6 kHz that correspond to the type-IImodes. The theory considers only growth rates and does not account for initialdisturbance levels.

An alternative to the approach used by Malik et al. (1994, 1996, 1999) ispresented by Koch et al. (2000), who found the nonlinear equilibrium solution ofthe primary flow. Koch et al. (2000) used the nonlinear equilibrium solution as areceptivity-independent basic state for a Floquet analysis of secondary instabilitiesof the saturated vortices. Yet another approach is by Janke & Balakumar (2000),who used an NPSE for the base flow and a Floquet analysis for the secondaryinstabilities. Both Koch et al. (2000) and Janke & Balakumar (2000) are in generalagreement with the various computations by Malik et al. (1994, 1996, 1999).

Hogberg & Henningson (1998) pursued a DNS approach to the problem of thestationary-vortex saturation and the ensuing secondary instability. These authorsimposed an artificial random disturbance at a point where the stationary vorticesare saturated. These disturbances enhance both the low- and high-frequency distur-bances downstream, and each frequency band has a distinct spatial location, withthe high-frequency disturbance located in the upper part of the boundary layerand the low-frequency disturbance located in the lower part. Spectral analysis ofthe resulting disturbance field shows that the most amplified high frequency issomewhat more than an order of magnitude higher frequency than the most ampli-fied traveling primary disturbance. Another high-frequency peak at approximatelytwice this frequency is also evident in the spectra. This peak likely corresponds toa type-II mode, although this feature is not described by the authors.

Wassermann & Klocker (2002) present another highly resolved DNS study ofnonlinear interactions of primary crossflow modes, their secondary instabilities,and eventual breakdown to turbulence. They emphasize disturbance wave packetsthat may be more realistic than single-mode disturbances. One of the most im-portant findings obtained from the wave-packet approach is that unevenly spacedprimary vortices of differing strengths can interact in such a way as to bring aboutan earlier onset of secondary instabilities and breakdown than would be found froma single-mode disturbance. Wassermann & Kloker (2002) also found that, whenthe forcing that initiates the high-frequency secondary instability in their simula-tion is removed, the secondary instability disturbances are convected downstream,out of the computational domain. This indicates that the secondary instability isconvective and that the explosiveness of the growth of the secondary instability is

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 21: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 433

not associated with an absolute instability. The advantage of the DNS solution byWassermann & Kloker (2002) is its ability to reveal the rather intimate details ofthe breakdown process. As such, their work is one of the foundation contributions.

To date, the various approaches to the secondary instability problem (experi-mental, NPSE, and DNS) have achieved rather remarkable agreement in terms ofidentifying the basic mechanisms of the secondary instability, unstable frequen-cies, mode shapes, and growth rates. A comparison of three of the most recentefforts is shown in Figure 5. This comparison shows agreement on the locationof the breakdown and indicates that it is associated with an inflection point in thespanwise direction (an extremum in∂U/∂z).

4. SWEPT CYLINDER

The discussion of the crossflow instability in the preceding section pertains toboundary layers with characteristics typical of large portions of swept wings overwhich boundary layers develop in response to weakly favorable pressure gradientand low wall and streamline curvature. In contrast, boundary layers in the leading-edge region of swept wings can exhibit different stability characteristics than thoseobserved farther downstream owing to the much stronger pressure gradient, wallcurvature, and streamline curvature that exist near the leading edge. These flowsare referred to as swept-cylinder flows to distinguish them from both swept-wingand attachment-line flows.

The swept-cylinder flow near the leading edge is distinct from the flow alongthe attachment line. Flow along the attachment line has stability characteristicsdifferent from swept-cylinder flow because of the various roles of surface andstreamline curvature in the two configurations. Attachment-line stability and con-tamination are discussed at length by Reed & Saric (1989) but are outside of thescope of this review. Reed et al. (1996) and Theofilis (1998) discuss recent workconcerning attachment-line issues.

For swept-cylinder flows, one of the principal recent findings is the discoveryby Itoh (1994) that highly curved inviscid streamlines produce a centrifugal-typeboundary-layer instability that is unique from the other 3-D boundary-layer insta-bilities that had been identified at the time of an earlier review by Reed & Saric(1989). Because the flow near the leading edge of swept cylinders features stronglycurved streamlines, it may exhibit a streamline-curvature instability similar to thatobserved at small Reynolds numbers on rotating disks. Although it has yet to bedemonstrated to have flight applicability, the streamline-curvature instability mayplay an important role under some set of conditions.

Itoh (1996) specifically examined the flow in the leading-edge region of sweptcylinders. The key findings are that, for intermediate values of the inclination ofthe inviscid streamline to the chord direction (i.e., starting somewhat downstreamof the leading edge), the streamline-curvature instability has the lowest criticalReynolds numbers of all instability modes and that the disturbances produced

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 22: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

434 SARIC ¥ REED ¥ WHITE

Figure 5 Mode-I velocity fluctuation contours: (a) Figure 7 from Malik et al.(1999), (b) Figure 20b from Wassermann & Kloker (2002), and (c) Figure 11from White & Saric (2003).

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 23: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 435

by the streamline-curvature instability take the form of streamwise vortices, andthat these vorices are even more closely aligned with the inviscid stream than arecrossflow vortices. Very close to the leading edge where the streamline inclinationis large, crossflow waves are expected to be the most unstable disturbances. For afixed value of the streamline inclination, linear-stability neutral curves for the sweptcylinder exhibit a two-lobed structure that suggests competition between crossflowand streamline-curvature instabilities, similar to what is observed for rotating-diskboundary layers. Itoh (1996) showed that the dynamics of this competition dependon sweep angle. As the sweep is increased the minimum critical Reynolds numberfor the streamline curvature instability is unchanged, but the point at which thecritical Reynolds number of the streamline-curvature instability is less than thatof the crossflow instability moves farther downstream, and this means that in thecases of high sweep, streamline curvature effects will be observed over a largerregion than is typical of lower-sweep configurations. However, Itoh (1996) pointedout that this result is only valid close to the leading edge, so the results shouldnot be taken to suggest that the streamline-curvature instability dominates fartherdownstream where the basic state deviates from a leading-edge-type flow.

Takagi & Itoh (1998) and Tokugawa et al. (1999) performed swept-cylinder ex-periments to verify the theoretical predictions by Itoh (1996) regarding the behaviorof the streamline-curvature instability in the leading-edge region. Controlled point-source disturbances are introduced, and a developing disturbance wedge providesa means of separating crossflow-related and streamline-curvature-related distur-bances. Itoh (1966) predicted that crossflow and streamline-curvature instabilitiespropagate in opposite directions, and Tokugawa et al. (1999) observed that, in-side the disturbance wedge, disturbances with phase-distributions characteristicof the crossflow and streamline-curvature instabilities dominate in two separatedisturbance-amplitude maxima that exist on either side of the wedge, providingconfirmation of the general findings by Itoh (1996). More recently, Itoh (2000)included nonparallel effects, and Yokokawa et al. (2000) used a DNS model. TheDNS results show that disturbances that propagate in the direction associated withthe streamline-curvature instability are characterized by vorticity oriented primar-ily in the wall-normal direction in contrast with the previous analyses by Itoh (1995,1996) that showed streamline-curvature disturbance modes are vortices orientedin the streamwise and not the wall-normal direction. This discrepancy has not beenresolved.

5. CONCLUSIONS

Boundary-layer transition in 3-D flows is a complicated process involving com-plex geometries, multiple instability mechanisms, and nonlinear interactions. Yetsignificant progress has recently been made toward understanding the instabilityand transition characteristics of 3-D flows. In terms of the crossflow problem, thepast decade has produced several important discoveries that include tools, such as

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 24: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

436 SARIC ¥ REED ¥ WHITE

■ instrumentation that can be applied to the flight-test environment,■ POD methods to interpret wind-tunnel and flight-test transition data,■ validation with careful experiments of NPSE codes to predict all aspects of

stationary disturbance growth.

Several important factors have also been identified:

■ environmental conditions on the appearance of stationary and traveling waves,■ secondary instability causing local transition in stationary-crossflow-domina-

ted flows,■ extreme sensitivity of the stationary disturbance to leading-edge, very small,

surface roughness,■ nonlinear effects and modal interaction,■ extreme sensitivity of stationary-wave growth to very weak convex curvature.

One result has been a better understanding of rotating-disk behavior. In addition,by carefully studying the basic physics, these advances have led to the promisingapplication of artificial roughness at the leading edge to control the crossflowinstability and delay transition on swept wings in flight.

The study of 3-D boundary-layer instability still offers challenges to the fluid-mechanics community. Important factors such as receptivity—the process bywhich external disturbances enter the boundary layer and create the initial condi-tions for an instability—are still not completely understood. Yet in spite of this,careful experiments and companion accurate computations have resulted in signif-icant progress toward understanding a difficult problem and offer promise of evenfurther advances in the future.

ACKNOWLEDGMENTS

This work was sponsored (in part) by the Air Force Office of Scientific Research,United States Air Force, under grant number F49620-97-1-0520.

The Annual Review of Fluid Mechanicsis online at http://fluid.annualreviews.org

LITERATURE CITED

Arnal D. 1994. Predictions based on linear the-ory.Progress in transition modeling, AGARDRep. No. 709

Arnal D, Casalis G, Reneaux J, Coustieux J.1997. Laminar-turbulent transition: researchand applications in France.AIAA Pap. No.97-1905

Arnal D, Michel R, eds. 1990.Laminar-Turbulent Transition, Vol. III.New York:Springer. 710 pp.

Balachandar S, Streett CL, Malik MR.1992. Secondary instability in rotating-disk flow. J. Fluid Mech. 242:323–47

Balakumar P, Malik MR. 1991. Waves pro-duced from a harmonic point source in a su-personic boundary layer.AIAA Pap. No. 91-1646

Bertolotti FP. 2000. Receptivity of three-dimensional boundary layers to localized

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 25: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 437

wall roughness and suction.Phys. Fluids12:1799–809

Bippes H. 1997. Environmental conditions andtransition prediction in 3-D boundary layers.AIAA Pap. No. 97-1906

Bippes H. 1999. Basic experiments on tran-sition in three-dimensional boundary lay-ers dominated by crossflow instability.Prog.Aerosp. Sci.35:363–412

Bippes H, Muller B, Wagner M. 1991. Mea-surements and stability calculations of thedisturbance growth in an unstable three-dimensional boundary layer.Phys. Fluids3:2371–77

Bippes H, Nitschke-Kowsky P. 1990. Exper-imental study of instability modes in athree-dimensional boundary layer.AIAA J.28:1758–63

Boiko AV, Kozlov VV, Syzrantsev VV, Shcher-bakov VA. 1995. Experimental investiga-tion of high-frequency secondary distur-bances in a swept-wing boundary layer.J. Appl. Mech. Tech. Phys.36:385–93

Boiko AV, Kozlov VV, Syzrantsev VV,Shcherbakov VA. 1999. Active control oversecondary instability in a three-dimensionalboundary layer. Thermophys. Aeromech.6:167–78

Buck GA, Takagi S. 1997. Observations of trav-eling disturbances from a point source in ro-tating disk flow.AIAA Pap. No. 97-2299

Buck GA, Takagi S, Tokugawa N. 2000. Experi-mental investigation of traveling disturbancemodes in the rotating disk boundary layer.AIAA Pap. No. 2000-2658

Chapman KL, Reibert MS, Saric WS, GlauserMN. 1998. Boundary-layer transition detec-tion and structure identification through sur-face shear-stress measurements.AIAA Pap.No. 98-0782

Choudhari M. 1994. Roughness-induced gener-ation of crossflow boundary vortices in three-dimensional boundary layers.Theor. Com-put. Fluid Dyn.6:1–30

Collis SS, Lele SK. 1999. Receptivity to sur-face roughness near a swept leading edge.J.Fluid Mech.380:141–68

Corke TC, Knasiak KF. 1998. Stationary trav-

elling cross-flow mode interactions on a ro-tating disk.J. Fluid Mech.355:285–315

Crouch JD. 1993. Receptivity of three-dimensional boundary layers.AIAA Pap. No.93-0074

Crouch JD. 1994. Theoretical studies on recep-tivity of boundary layers.AIAA Pap. No. 94-2224

Crouch JD. 1995. Distributed excitation ofcross-flow vortices in three-dimensionalboundary layers. see Kobayashi 1995, pp.499–506

Crouch JD. 1997. Transition prediction andcontrol for airplane applications.AIAA Pap.No. 97-1907

Crouch JD, Crouch IWM, Ng LL. 2001. Es-timating the laminar/turbulent transition lo-cation in three-dimensional boundary layersfor CFD applications.AIAA Pap. No. 2001-2989

Crouch JD, Ng LL. 2000. Variable N-factormethod for transition prediction in three-dimensional boundary layers.AIAA J. 38:211–16

Dagenhart JR, Saric WS, Hoos JA, MousseuxMC. 1990. Experiments on swept-wingboundary layers. See Arnal & Michel 1990,pp. 369–80

Dagenhart JR, Saric WS, Mousseux MC, StackJP. 1989. Crossflow vortex instability andtransition on a 45-degree swept wing.AIAAPap. No. 89-1892

Deyhle H, Bippes H. 1996. Disturbancegrowth in an unstable three-dimensionalboundary layer and its dependence on ini-tial conditions. J. Fluid Mech. 316:73–113

Deyhle H, Hohler G, Bippes H. 1993. Experi-mental investigation of instability wave prop-agation in a 3-D boundary-layer flow.AIAAJ. 31:637–45

Duck PW, Hall P, eds. 1996.IUTAM Symposiumon Nonlinear Instability and Transition inThree-Dimensional Boundary Layers. Am-sterdam: Kluwer. 436 pp.

Faller AJ. 1991. Instability and transition ofdisturbed flow over a rotating disk.J. FluidMech.230:245–70

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 26: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

438 SARIC ¥ REED ¥ WHITE

Fasel HF, Saric WS, eds. 2000.Laminar-Turbulent Transition V. Berlin: SpringerVerlag. 686 pp.

Fedorov AV. 1989. Excitation of waves of in-stability of secondary flow in the boundarylayer on a swept wing.J. Appl. Mech. Tech.Phys.29:643–48

Hall P, Malik MR, Poll DIA. 1984. On thestability of an infinite swept attachment-line boundary layer.Philos. Trans. R. Soc.London Ser. A395:229–45

Haynes TS, Reed HL. 2000. Simulation ofswept-wing vortices using nonlinear parab-olized stability equations.J. Fluid Mech.405:325–49

Herbert T. 1997a. Parabolized stability equa-tions.Annu. Rev. Fluid Mech.29:245–83

Herbert T. 1997b. Transition prediction andcontrol for airplane applications.AIAA Pap.No. 97-1908

Herbert T, Lin N. 1993. Studies of boundary-layer receptivity with parabolized stabilityequations.AIAA Pap. No. 93-3053

Hogberg M, Henningson D. 1998. Secondaryinstability of crossflow vortices in Falkner-Skan-Cooke boundary layers.J. Fluid Mech.368:339–57

Huerre P, Monkewitz PA. 1990. Local andglobal instabilities in spatially developingflows.Annu. Rev. Fluid Mech.22:473–537

Itoh N. 1985. Stability calculations of the three-dimensional boundary-layer flow on a rotat-ing disk. See Kozlov 1985, pp. 463–70

Itoh N. 1994. Instability of three-dimensionalboundary layers due to streamline curvature.Fluid Dyn. Res.14:353–66

Itoh N. 1995. Effects of wall and streamline cur-vatures on instability of 3-D boundary layers.See Koyabashi 1995, pp. 323–30

Itoh N. 1996. Simple cases of the streamline-curvature instability in three-dimensionalboundary layers.J. Fluid Mech.317:129–54

Itoh N. 1997. Development of an axisymmetricwavepacket in rotating-disk flow.Fluid Dyn.Res.21:87–100

Itoh N. 1998. Theoretical description of insta-bility waves in the flow on a rotating disk.Part 1. False-expansion method applied to

linear stability equations.Trans. Jpn. Soc.Aerosp. Sci.40:262–79

Itoh N. 2000. Non-parallel stability analysisof three-dimensional boundary layers alongan infinite attachment line.Fluid Dyn. Res.27:143–61

Janke E. 2001. Receptivity and transition con-trol of swept-wing boundary-layers: effectsof surface curvature and nonlinearity.AIAAPap. No. 2001-2980

Janke E, Balakumar P. 2000. On the secondaryinstability of three-dimensional boundarylayers.Theor. Comput. Fluid Dyn.14:167–94

Jarre S, Le Gal P, Chauve MP. 1996a. Experi-mental study of rotating disk flow instability.I. Natural flow.Phys. Fluids8:496–508

Jarre S, Le Gal P, Chauve MP. 1996b. Experi-mental study of rotating disk flow instability.II. Forced flow.Phys. Fluids8:2985–94

Juillen JC, Arnal D. 1990. Etude exp´erimentaledu de clenchement de la transition par ru-gosites et par rainer sur de bord d’attaqued’une aile en fleche en ´encoulement in-compressible.Rapp. Final No. 51/5018.35.CERT/ONERA, Toulouse

Kachanov YS. 1996. Experimental studies ofthree-dimensional instability of boundarylayer.AIAA Pap. No. 96-1976

Kachanov YS, Tararykin OI. 1990. The experi-mental investigation of stability and receptiv-ity of a swept-wing flow. See Arnal & Michel1990, pp. 499–509

Kawakami M, Kohama Y, Okutsu M. 1999.Stability characteristics of stationary cross-flow vortices in three-dimensional boundarylayer.AIAA Pap. No. 99-0811

Kobayashi R, ed. 1995.Laminar-TurbulentTransition, Vol. IV. Berlin: Springer Verlag.532 pp.

Koch W, Bertolotti FP, Stolte A, Hein S. 2000.Nonlinear equilibrium solutions in a three-dimensional boundary layer and their sec-ondary instability.J. Fluid Mech.406:131–74

Kohama Y, Onodera T, Egami Y. 1996. Designand control of crossflow instability field. SeeDuck & Hall 1996, pp. 147–56

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 27: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

3-D BOUNDARY LAYERS 439

Kohama Y, Saric WS, Hoos JA. 1991.A high-frequency, secondary instability ofcrossflow vortices that leads to transition.Proc. R. Aerosp. Soc. Conf. Bound.-LayerTransit. Control Conf.Peterhouse College,Cambridge, UK pp. 4.1–4.13

Kozlov VV, ed. 1985.Laminar-Turbulent Tran-sition, Vol. II. Berlin: Springer Verlag. 759pp.

Lerche T. 1996. Experimental investigation ofnonlinear wave interactions and secondaryinstability in three-dimensional boundary-layer flow.Eur. Turbul. Conf., 6th, Lausanne,ed. S. Gavrilakis, L Machiels, PA Monke-witz, pp. 357–60

Lin RS, Reed HL. 1993. Effect of curva-ture on stationary crossflow instabilities ofa three-dimensional boundary layer.AIAA J.31:1611–17

Lingwood RJ. 1995. Absolute instability of theboundary layer on a rotating disk.J. FluidMech.299:17–33

Lingwood RJ. 1996. An experimental studyof absolute instability of the rotating diskboundary layer flow.J. Fluid Mech.314:375–405

Mack LM. 1984. Boundary-layer linear stabil-ity theory.Spec. Course Stab. Transit. Lami-nar Flows, AGARD Rep. 709, pp. 3.1–81

Malik MR. 1986. The neutral curve for station-ary disturbances in rotating disk flow.J. FluidMech.164:275

Malik MR, Li F, Chang CL. 1994. Crossflowdisturbances in three-dimensional boundarylayers: nonlinear development, wave interac-tion and secondary instability.J. Fluid Mech.268:1–36

Malik MR, Li F, Chang CL. 1996. Nonlinearcrossflow disturbances and secondary insta-bilities in swept-wing boundary layers. SeeDuck & Hall 1996, pp. 257–66

Malik MR, Li F, Choudhari MM, Chang CL.1999. Secondary instability of crossflow vor-tices and swept-wing boundary layer transi-tion. J. Fluid Mech.399:85–115

Meyer F, Kleiser L. 1990. Numerical simula-tion of transition due to crossflow instability.See Arnal & Michel 1990, pp. 609–19

Muller B. 1990. Experimental study of the trav-eling waves in a three-dimensional boundarylayer. See Arnal & Michel 1990, pp. 489–99

Muller B, Bippes H. 1989. Experimental studyof instability modes in a three-dimensionalboundary layer.Fluid Dyn. Three-Dimens.Turbul. Shear Flows Transit., AGARD CP438, pp. 13.1–15

Nitschke-Kowsky P, Bippes H. 1988. Instabilityand transition of a three-dimensional bound-ary layer on a swept flat plate.Phys. Fluids31:786–95

Ng LL, Crouch JD. 1999. Roughness-inducedreceptivity to crossflow vortices on a sweptwing. Phys. Fluids11:432–38

Poll DIA. 1985. Some observations of the tran-sition process on the windward face of a longyawed cylinder.J. Fluid Mech. 150:329–56

Radeztsky RH Jr, Reibert MS, Saric WS. 1994.Development of stationary crossflow vorticeson a swept wing.AIAA Pap. No. 94-2373

Radeztsky RH Jr, Reibert MS, Saric WS. 1999.Effect of isolated micron-sized roughnesson transition in swept-wing flows.AIAA J.37:1371–77

Reed HL, Lin RS. 1987. Stability of three-dimensional boundary layers.SAE Pap. No.87-1857

Reed HL, Saric WS. 1989. Stability of three-dimensional boundary layers.Annu. Rev.Fluid Mech.21:235–84

Reed HL, Saric WS, Arnal D. 1996. Linearstability theory applied to boundary layers.Annu. Rev. Fluid Mech.28:389–428

Reibert MS, Saric WS, Carrillo RB Jr, Chap-man KL. 1996. Experiments in nonlinear sat-uration of stationary crossflow vortices in aswept-wing boundary layer.AIAA Pap. No.96-0184

Reshotko E. 1976. Boundary-layer stability andtransition.Annu. Rev. Fluid Mech.8:311–49

Reshotko E. 1994. Boundary-layer instability,transition and control.AIAA Pap. No. 94-0001

Saric WS. 1994. G¨ortler vortices.Annu. Rev.Fluid Mech.26:379–409

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 28: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

22 Nov 2002 11:40 AR AR159-FM35-17.tex AR159-FM35-17.sgm LaTeX2e(2002/01/18)P1: GLM

440 SARIC ¥ REED ¥ WHITE

Saric WS, Carrillo RB Jr, Reibert MS.1998. Nonlinear stability and transition in3-D boundary layers.Meccanica33:469–87

Saric WS, Reed HL. 2002. Supersonic laminarflow control on swept wings using distributedroughness.AIAA Pap. No. 2002-0147

Saric WS, Reed HL, Kerschen EJ. 2002.Boundary-layer receptivity to freestream dis-turbances.Annu. Rev. Fluid Mech.34:291–319

Saric WS, Yeates LG. 1985. Generation ofcrossflow vortices in a three-dimensionalflat-plate flow. See Kozlov VV 1985, pp.429–37

Spalart PR. 1993. Numerical study of transi-tion induced by suction devices. InNear-WallTurbulent Flows, pp. 849–58. The Nether-lands: Elsevier

Takagi S, Itoh N. 1998. Dispersive evolution ofcross-flow disturbances excited by an airjetcolumn in 3-D boundary layer.Fluid Dyn.Res.22:25–42

Takagi S, Itoh N, Tokugawa N. 1998. Char-acteristics of unsteady disturbances due tostreamline-curvature instability in a rotating-disk flow.AIAA Pap. No. 98-0341

Takagi S, Itoh N, Tokugawa N, Nishizawa A.2000. Characteristic features of traveling dis-turbances originating from a point source inrotating-disk flow. See Fasel & Saric 2000,pp. 637–42

Theofilis V. 1998. On linear and nonlin-ear instability of the incompressible sweptattachment-line boundary layer.J. FluidMech.355:193–227

Tokugawa N, Takagi S, Itoh N. 1999. Ex-citation of streamline-curvature instabilityin three-dimensional boundary layer on ayawed cylinder.AIAA Pap. No. 99-0814

Vonderwell MP, Riahi DN. 1993. Weakly non-linear development of rotating disk flow.Int.J. Eng. Sci.31:549–58

von Doenhoff AE, Braslow AL. 1961. The ef-fect of distributed roughness on laminar flow.In Boundary-Layer Control, ed. G. Lach-mann, 2:657–81. New York: Pergamon

Wang M, Herbert T, Stuckert GK. 1994.Crossflow-induced transition in compress-ible swept-wing flows.AIAA Pap. No. 94-2374

Wassermann P, Kloker M. 2002. Mechanismsand control of crossflow-vortex induced tran-sition in a 3-D boundary layer.J. Fluid Mech.456:49–84

White EB, Saric WS. 2003. Secondary insta-bility of crossflow vortices.J. Fluid Mech.Submitted

White EB, Saric WS, Gladden RD, Gabet PM.2001. Stages of swept-wing transition.AIAAPap. No. 2001-0271

Wintergerste T, Kleiser L. 1995. Direct nu-merical simulation of transition in a three-dimensional boundary layer. In Transi-tional Boundary Layers in Aeronautics, ed.R Henkes, J van Ingen, pp. 145–54. Amster-dam/New York: North-Holland

Yokokawa Y, Fukunishi Y, Itoh N. 2000. Nu-merical study of excitation of two differentinstabilities in three-dimensional boundarylayer on a yawed cylinder. See Fasel & Saric2000, pp. 613–18

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 29: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

29 Nov 2002 13:42 AR AR159-17-COLOR.tex AR159-17-COLOR.SGM LaTeX2e(2002/01/18)P1: GDL

Fig

ure

2Ve

loci

tyco

ntou

rsfr

omR

eibe

rtet

al.(

1996

).

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 30: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

29 Nov 2002 13:42 AR AR159-17-COLOR.tex AR159-17-COLOR.SGM LaTeX2e(2002/01/18)P1: GDL

Fig

ure

4Ve

loci

tyco

ntou

rco

mpa

rison

s.

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 31: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

P1: FDS

November 22, 2002 11:16 Annual Reviews AR159-FM

Annual Review of Fluid MechanicsVolume 35, 2003

CONTENTS

STANLEY CORRSIN: 1920–1986, John L. Lumley and Stephen H. Davis 1

AIRCRAFT ICING, Tuncer Cebeci and Fassi Kafyeke 11

WATER-WAVE IMPACT ON WALLS, D. H. Peregrine 23

MECHANISMS ON TRANSVERSE MOTIONS IN TURBULENT WALLFLOWS, G. E. Karniadakis and Kwing-So Choi 45

INSTABILITIES IN FLUIDIZED BEDS, Sankaran Sundaresan 63

AERODYNAMICS OF SMALL VEHICLES, Thomas J. Muellerand James D. DeLaurier 89

MATERIAL INSTABILITY IN COMPLEX FLUIDS, J. D. Goddard 113

MIXING EFFICIENCY IN STRATIFIED SHEAR FLOWS, W. R. Peltierand C. P. Caulfield 135

THE FLOW OF HUMAN CROWDS, Roger L. Hughes 169

PARTICLE-TURBULENCE INTERACTIONS IN ATMOSPHERIC CLOUDS,Raymond A. Shaw 183

LOW-DIMENSIONAL MODELING AND NUMERICAL SIMULATIONOF TRANSITION IN SIMPLE SHEAR FLOWS, Dietmar Rempfer 229

RAPID GRANULAR FLOWS, Isaac Goldhirsch 267

BIFURCATING AND BLOOMING JETS, W. C. Reynolds, D. E. Parekh,P. J. D. Juvet, and M. J. D. Lee 295

TEXTBOOK MULTIGRID EFFICIENCY FOR FLUID SIMULATIONS,James L. Thomas, Boris Diskin, and Achi Brandt 317

LEVEL SET METHODS FOR FLUID INTERFACES, J. A. Sethianand Peter Smereka 341

SMALL-SCALE HYDRODYNAMICS IN LAKES, Alfred Wuestand Andreas Lorke 373

STABILITY AND TRANSITION OF THREE-DIMENSIONAL BOUNDARYLAYERS, William S. Saric, Helen L. Reed, Edward B. White 413

SHELL MODELS OF ENERGY CASCADE IN TURBULENCE, Luca Biferale 441

FLOW AND DISPERSION IN URBAN AREAS, R. E. Britter and S. R. Hanna 469

vii

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.

Page 32: Personal Homepages for the University of Bath - STABILITY AND …staff.bath.ac.uk/ensdasr/PAPERS/ReedSaric.pdf · 2010. 5. 19. · stability neutral curve has two lobes: one for the

P1: FDS

November 22, 2002 11:16 Annual Reviews AR159-FM

viii CONTENTS

INDEXESSubject Index 497Cumulative Index of Contributing Authors, Volumes 1–35 521Cumulative Index of Chapter Titles, Volumes 1–35 528

ERRATAAn online log of corrections to Annual Review of Fluid Mechanics chaptersmay be found at http://fluid.annualreviews.org/errata.shtml

Ann

u. R

ev. F

luid

Mec

h. 2

003.

35:4

13-4

40. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F B

AT

H o

n 05

/19/

10. F

or p

erso

nal u

se o

nly.