You are on page 1of 80

CHAPTER 1

1. INTRODUCTION

HISTORY OF VISCOUS FLUID DYNAMICS

In physics and engineering , Fluid dynamicsis a subdiscipline of fluid

dynamics that describes the flow of fluids (liquids and gases). It has several

subdisciplines, including aerodynamics (the study of air and other gases in motion)

and hydrodynamics (the study of liquids in motion). Fluid dynamics has a wide

range of applications, including calculating forces and moments on aircraft,

determining the massflow rate of petroleum through pipelines, predicting

weatherpatterns, understanding nebulae in interstellar space and modelling fission

weapon detonation.

Fluid dynamics offers a systematic structurewhich underlies these practical

disciplinesthat embraces empirical and semi-empirical laws derived from flow

measurement and used to solve practical problems. The solution to a fluid dynamics

problem typically involves the calculation of various properties of the fluid, such

as flow velocity, pressure, density, and temperature, as functions of space and time.

Before the twentieth century, hydrodynamics was synonymous with fluid

dynamics. This is still reflected in names of some fluid dynamics topics,

like magnetohydrodynamics and hydrodynamic stability, both of which can also be

applied to gases.
CHAPTER 2

PRELIMINARIES

Definition: Reynolds Number

The dimensionless quantity Re defined as

UL UL
Re = = ,
v

Where, U, L, and are some characteristics values of the velocity, length, density

and viscosity of the fluid respectively is known as the Reynolds number.

Definition: Turbulent Flow

The type of flow in which fluid particles do not move in layers

or lamination is called turbulent flow.

Definition: Turbulence

Turbulence is three-dimensional time dependent motion in which vortex

stretching causes velocity fluctuations to spread to all wavelengths between a

minimum determined by viscous forces and maximum determined by boundary

condition of the flow. It is the usual state of fluid motion except at low Reynolds

numbers.
Definition: Laminar Flow

The type of flow in which Fluid particles move in smooth layers or

in lamination is called laminar flow. If Reynolds number is less than or equal to

2000, the Flow will be laminar.

Definition: Isotropic

Statistically independent of direction

Example: u v w .

In practice isotropic turbulence must also be homogeneous.

Definition: Homogeneous

Statistically independent of position in space.

Example: independent of x .

Not necessarily isotropic.

Definition: Kolmogorov length scales

Smallest scales in the spectrum that form the viscous sub-layer range. In this

range, the energy input from nonlinear interactions and the energy drain from viscous

dissipation are in exact balance. The small scales have high frequency, causing

turbulence to be locally isotropic and homogeneous.

Definition: Lyapunov exponent

The Lyapunov exponent of a dynamical system provides a time averaged

measure of the rate of divergence of two trajectories that were initially nearby.
Definition: Eddy viscosity

Eddy viscosity is, formally, the constant of proportionality between turbulent

(Reynolds) stresses and mean (large-scale) strain rate, analogous to physical viscosity

in Newtons law of viscosity.

Definition: Kolmogorov scale

Kolmogorov scale is another name for dissipation scales. These scales were

predicted on the basis of dimensional analysis as part of the K41 theory.

Definition: RANS

RANS is the acronym for Reynolds-Averaged NavierStokes approaches to

turbulence calculation. As will be evident in later lectures, RANS approaches are

essentially the opposite of DNS, namely, nearly all scales of the solution must be

modeled under a RANS formalism; only (time) mean quantities are directly

computed.

Definition: Reynolds stress

Reynolds stresses arise from the averaging procedure used in deriving the

RANS equations. They are the additional unknowns that create the closure problem.

Details of the derivation of the RANS equations will be given in Chap.-2 where

modelling of the Reynolds stresses will also be discussed.


Definition: Transition

Transition refers to a change from one qualitative state of flow to another, and

in particular in the context of turbulence studies, the transitions from laminar to

turbulent.

Definition: The maximal Lyapunov exponent

The maximal Lyapunov exponent can be defined as follows:

1 |()|
= lim lim
0 0 |0 |

The limit 0 0 ensures the validity of the linear approximation at any time.

For discrete time system (maps or fixed point iterations) +1 = ( ) , for an

orbit starting with x0 this translates into:

1
1
(0 ) = lim | ( )|

=0

Definition: The Lyapunov spectrum

For a dynamical system with evolution equation in an ndimensional phase

space, the spectrum of Lyapunov exponents {1, 2, 3 n} in general,

depends on the starting point x0. However, we will usually be interested in

the attractor (or attractors) of a dynamical system, and there will normally be one set

of exponents associated with each attractor. The choice of starting point may

determine which attractor the system ends up on, if there is more than one. Note:

Hamiltonian systems do not have attractors, so this particular discussion does not
apply to them.) The Lyapunov exponents describe the behaviuor of vectors in the

tangent space of the phase space and are defined from the Jacobian matrix

(
()
) = |

The matrix describes how a small change at the point x0 propagates to the

final point ( ). The limit


( ) = ( . ( ))

defines L(x0) a matrix (the conditions for the existence of the limit are given by

the Oseledec theorem). If i(x0) are the eigenvalues L(x0) of , then the Lyapunov

exponents i are defined by

(0 ) = log (0 )

The set of Lyapunov exponents will be the same for almost all starting points

of an ergodic component of the dynamical system.

Definition: Couette Flow

The plane couette flow or simple shear flow is the flow between two parallel

plates one of which is at rest and the other moving with a uniform velocity U in its

own plane and the pressure gradient is taken to be zero.

Definition: Poiseuille Flow

In plane poiseullie flow the pressure gradient is non-zero, but the plates are

kept stationary.
CHAPTER 3

PREDICTION OF TURBULENT TRANSITION IN BOUNDARY


LAYERS USING THE TURBULENT POTENTIAL MODEL

1. INTRODUCTION

Predicting the onset of turbulent flow is a critical component of many

engineering and environmental flows. The characteristics of laminar and turbulent

boundary layers are so different that the precise location of this relatively abrupt

transition can have a profound influence on the overall drag, heat transfer, and

performance properties of devices that operate in the transitional regime. The

prediction of boundary layer transition is complicated by the fact that it does not

correspond very directly to the onset of instability. Stability analysis for boundary

layers is well developed and very predictive of the behaviour of small disturbances.

However, the instabilities go through a series of complex nonlinear and three-

dimensional processes before turbulence itself actually develops. In addition, stability

analysis is less helpful with the prediction of bypass transition where external free-

stream disturbances bypass the classic instability mechanisms and initiate the

nonlinear three-dimensional transition process directly.

Traditional methods for predicting transition rely on correlations. For example,

[20] suggest the implicit correlation

x,tr 0.312(m 0.11)


1690Re-1/2 -0.528
4.8 99
2
Re1/2
x,tr T
2

where Rex,tr is the transition Reynolds number based on the local free-stream

velocity U, the parameter m is related to the dimensionless pressure gradient

( m = 0 for zero pressure gradient), 99 is the local 99% boundary layer thickness,
and T (2/3)k / U is the local free-stream turbulence intensity. This is only one

of many proposed correlations. While they appear relatively simple, they are

actuallyquite awkward to implement in a general purpose CFD code. In a general

code, correlations require each point on the boundary to determine non-local

quantities, such as free-stream values, distance downstream from the leading edge

and boundary layer thickness. Unambiguous definitions for such quantities in a

general situation are very difficult to formulate. Many CFD codes simply resort to

requiring the user to specify the transition location. Once the transition location has

been determined, there is the further difficulty of determining how the turbulence

model should be prompted to become active. Solutions to this problem range from

specifying large turbulent kinetic energy at the wall where transition is expected to

occur to specifying local large source terms in the turbulence evolution equations

(artificial production). None of these methods of tripping the turbulence model are

particularly reflective of the actual physical transition process, and add further

ambiguity to the already uncertain transition location. In addition, tripping of this

sort is largely code dependent, so that the same theoretical model can produce

different results in different codes, or even within the same code using different

meshes.

An alternative approach is to use the turbulence model itself to predict

the transition location. This is a very natural approach in by pass transition, since the

model is effectively on in the free stream anyway to predict the free-stream

turbulence. A number of studies of boundary layer bypass transition prediction using

low Reynolds number k models have been performed, and [13] gave good reviews

of how various flavours of the k models perform. The overall conclusion is that

none performs well for all the flows considered. More recent work usinga k
intermittency model (Suzen and Huang 2001) has shown more success. However,

none of these models attempts to predict natural transition or relaminarization.

2. MOTIVATION

While the idea of using Reynolds averaged NavierStokes (RANS) models to

predict transitionmay have originally been motivated by practical considerations such

as ease of implementation,it also has a solid theoretical justification that has not been

discussed previously. The derivationof the RANS and accompanying Reynolds stress

transport (RST) equations are simply amathematical reformulation of the governing

NavierStokes equations. There are no physicalassumptions or restrictions on the

range of applicability of these equations, as presented below.


+ . () = p + . vu . R (1 a)

. u = 0 (1 b)



2

+ . R = v R (

+
)+ ( + )





2v
( + + )
(2)

whereuis the mean velocity, and Ris the single point, second-order correlation of the

fluctuatingvelocity. The derivation of these equations can be found in numerous

textbooks. The first termin parentheses is the production term. This term does not

require a model if equation (2) issolved. The subsequent terms are pressure strain,

dissipation, turbulence transport and pressuretransport terms, all of which require a

model.

This rearrangement of the equations is undertaken in order to isolate what is

computable(the mean flow), from the underlying (incomputable) turbulent velocity


fluctuations. The pricethat is paid for isolating this computable system from the

governing (and essentially intractable)NavierStokes equations is that the equations

are unclosed, and models must be constructed torepresent the effect of the turbulence

fluctuations on the mean flow. It is these models, notthe equations themselves, that

may, or may not, be able to capture the nonlinear transitionto turbulence. The RST

equation (2) is the basis for many turbulence models (including k/,algebraic

Reynolds stress, and turbulent potential models) and is the starting point for

thisdiscussion.

In a fully turbulent flow all the unclosed terms in the RST equations require

modelling,none can be neglected. However, in a transitioning flow, the fluctuations

are relatively smalland one might expect the turbulent transport term, which involves

a velocity triple product, tobe very small. Linear stability analysis is akin to assuming

that the turbulent transport term isidentically zero. Furthermore, this term can only

redistribute energy spatially. It does not causeany increase in overall turbulence

levelsthe hallmark of a transitioning flow. The pressuretransportterm (last term in

equation (2)) is small in fully developed turbulent boundary layers,and also can only

redistribute energy, so we will assume that like the turbulent transport term itis not a

critical process in transition. The dissipation term (involving viscosity) is also

expectedto be small when the transition process initiates. Eventually, we know that

the turbulence willcreate very strong local gradients so that the dissipation term will

balance the other terms inthe RST equation. However, early in the transition process,

the gradients will be relativelyweak and the turbulence will grow very rapidly,

indicating dissipation is not dominant. Thedissipation model is expected to have a

significant influence on the exact nature of the overshoot(in turbulence intensities and

skin friction) that occurs at the end of transition just before fullydeveloped turbulence
is obtained, but it is unlikely to have much effect on the much more criticaltransition

location.

The final modelled term, the pressurestrain correlation cannot be neglected in

thetransition process. The equation for the fluctuating pressure

= 2 , ,

+ , ( ) (3)
,

shows that the pressure has two components, one of which is dependent on the mean

shear,and is linearly proportional to the fluctuating velocity. This rapid pressure

strain term is of thesame order of magnitude as the production term, and tends to

reduce some, but not all, of theproduction. Rapid distortion theory (RDT) can tell us a

great deal about how this term shouldbehave in certain limits. The RDT analysis is

formally restricted to spatially homogeneousfluctuations but is otherwise well suited

to the problem of transition.

In summary, turbulence models that accurately capture production and rapid

pressurestrain processes are expected to have a good chance of predicting turbulent

transition. Modelsfor the pressure transport, turbulent transport, and dissipation are

only expected to influencethe very last stages of transition (the overshoot recovery)

and should have little effect on thecritical transition location. Most models for the

rapid pressure strain usually only capturerapid shear correctly because this is the

important case for boundary layer and free-shear flows.Reference [12] have a

modelling framework that can capture any rapidly distorting flow.

Due to the complexity of solving the RST equations, two-equation models

such as k/ arevery popular. The k-equation is derived by taking one half of the trace

of equation (2):

2
1
1 (4)

+ . = { ( + )} ( ) + 2

2 2

Two very important changes take place when only the k-equation is solved.

Firstly, pressurestrain disappears entirely. This is because this term only redistributes

energy between variousReynolds stress components, it does not influence the overall

energy content. Unlike spatialredistribution (transport terms) this intercomponent

redistribution is important, particularly inthe boundary layer. It takes energy out of the

amplified streamwise fluctuations and transfersit to the normal and spanwise

fluctuations, damping the streamwise fluctuation growth andfeeding the normal and

spanwise velocity fluctuations. Two-equation models are fundamentallyincapable of

capturing physical processes that involve intercomponent energy redistribution.

Inaddition, the production term is no longer exact. The Reynolds stress tensor in this

term mustnow be modelled. The classic Boussinesqeddy viscosity model is

2
= ( + ) (5)
3

where the eddy viscosity is given by = 2/. More complicated versions of this

equationare possible and are referred to as nonlinear eddy viscosity models or

algebraic Reynolds stressmodels. They all have the same fundamental problem for

predicting transition. These modelsassume that the fluctuations are somehow in

equilibrium with the mean flow. They suppose thata given shear level results in a

given turbulence level. The transition process is as far from anequilibrium situation as

possible. During transition, fluctuation amplitudes grow exponentiallyin time (moving

with the fluid), while shear levels remain almost constant.

Despite these difficulties, it is interesting to note that two-equation models still

can producequalitatively transition-like behaviour, with a rapid increase in turbulence


levels and skin frictionfrom low initial levels. This suggests that these differential

equations, and by inference themodelled RST equations, have the mathematical

capability to produce exponentially growingsolutions, and transitional behaviour.

With the added accuracy in the production and rapidpressurestrain terms, it is

expected that RST models could potentially constitute a veryaccurate transition

prediction methodology. However, classic RST models represent a

significantinvestment in programming complexity and computational resources. The

turbulent potentialmodel used in this work [9, 11, 21] represents an effective

compromise. It is a reformulation ofthe RST equations that retains the non-

equilibrium and energy redistribution physics, but whichcan be implemented at a

computational cost and programming complexity comparable to thevery popular two-

equation models. The turbulent potential models formulation and success

inpredicting fully turbulent flows has been discussed in prior publications [2, 10, 16].

In this paper,we focus on its unique abilities to predict transition in boundary layer

flows. We demonstratethe ability to predict bypass transition, the effects of pressure

gradients, natural transition,relaminarization, and the effect of noise levels on natural

transition.

3. MODEL EQUATIONS

The turbulent potential model used in this work is a slight modification of the

model proposedby [9]. The basic motivation and derivation remains the same.

Dk
= ( +
k) + P (6a)
Dt

D
) + (C1 P C2 )
= ( + (6b)
Dt k


= ( + ) + +

1/2
+ (1 1.5 ) ( ) (6c)


D 1/2
= ( + t) + + + + C ( ) k
Dt
(6d)

The constant and parameters are given by

k 1
P = , t = C , =
1 + 1.5

=
10| 1/2 |
[1 + ]

t 3 6
C = 0.21, Cp1 = 2.0 , Cp2 = , Cp4 = , C = 0.0033
t + 5 5 7
P
= 0.33 + 0.67

P
= 0.33 + 0.5

= 0.33

k2
Ce1 = 1.45, Ce2 = 1.83 0.16 exp (0.25 )

The slow pressure-strain models are



) ( ) k + Cp1 (2 1)
= ( +
k k


(k) (k) P
+ (Cp2 + Cp4 ) [ 25
]
/k (1 + Re ) k
and


= 2( + ) ( ) k Cp1 (1 )
k k

The fast pressure-strain models are



P
= Cp2
k

and

P P
= (1 Cp2 ) + Cp2 + C (2 1) 2
k k

This dissipation terms are given by


1 1
)2 2 + 2
= 2( +
k
and

=
k

Transport terms are modelled by the in the diffusion term. The two terms involving

in the turbulent potential equations (6c) and (6d) are the critical terms for

transition. Interestingly,because is so small, these terms have no influence on fully

developed turbulent flows. The model shows no perceptible difference in its

predictions when is set to zero in fully turbulent flows. However, it has a very

direct effect on transition, with higher values of causing transition to occur earlier.

We have used a value of 0.0033 in all the results presented in the paper, but it should

be observed that the model has the unique ability to explicitly control transition if the

user so desires.

The equations are a fairly standard implementation with slightly modified

definitions ofsome of the constants. Standard definitions for these constants will also

work quite adequately,these are just our own preferences. What is important to
emphasize is that although we aresolving transport equations this model is

nothing like a /model. In this model thepotentials are defined by the exact

relationship R = + , so k and are only usedto model the source terms in

the turbulent potential evolution equations. They are not used todirectly calculate the

Reynolds stress tensor, or the effect of the turbulence on the mean flow.

4. COMPUTATION RESULTS AND DISCUSSION

A number of different transitional flows have been calculated using the

present model, and arecompared to experimental data. The results presented below

include:transition in flat plateBlasius boundary layers with different levels of free-

stream turbulence intensity, including verylow level leading to natural transition, the

effects of variable pressure gradients and noise on flatplate transition, and

relaminarization in turbulent channel flow.

4.1. Flat plate boundary layer in zero pressure gradient:

The process of transition is studied by looking at the evolution of the friction

coefficient alongthe streamwise direction. The friction coefficient is a very sensitive

indicator of transition thatincreases dramatically as transition occurs. The model

predictions are compared to experimentaldata with different turbulence intensities.

The mean velocity is initially uniform flow for all casesand the initial values of

velocity U0 , turbulence Reynolds number ReT = k 2 /(), turbulenceintensity level

= (2/3)1/2 / for five experimental cases with different turbulenceintensities

are given in table 1. The initial values of the turbulent kinetic energy k are

determinedusing k = 3/2(TuU0 )2 . For the Tu = 0.03, 1.25 and 1.3% cases, the

initial turbulent dissipationrate is calculated from ReT using = k 2 /(ReT ) and the

value of ReT is assumed. The resultsare not very sensitive to reasonable values of ReT .

For the Tu = 3%(T3A) and Tu = 6%(T3B)cases, the data for k are available, and
dk
the initial values of are determined from = U ,where the subscript
dx

represents parameters in the free stream. The values of ReT for thesetwo cases listed in

table1 are just calculated from . The initial potentials and are set as2/3and zero

respectively. All experiments were performed in air so = 1.55 105 was usedin

every case.

The friction coefficient is plotted against,Rex , the Reynolds number based on

downstreamposition, in figure-1. The friction coefficient correlations for laminar and

1/2 1/7
turbulence flows,C = 0.664 Re and 0.027 respectively, are also plotted

with dashed curves forcomparison. Overall, the ability of the present model to predict

the critical transition location isquite good. However, the natural transition case, Tu =

0.03%, shows some discrepancy with the

Table 1.

Initial flow parameters at the leading edge of a zero pressure gradientboundary

layer.

( ) (%) Source

24.4 100 0.03 [14]

22 250 1.25 [1]

14.42 250 1.3 [5]

5.4 200 3.0 ERCOFTAC, T3A

9.4 200 6.0 ERCOFTAC, T3B


Figure 1:

Transition in zero pressure gradient boundary layer at various initialturbulence

intensities. The symbols represent experiment data and the curvesare the present

results. The dashed curves are fully laminar and fully turbulentcorrelations.

experimental data of [14] at 2.8 106. Actually, this discrepancy is somewhat

expected giventhe model initial conditions. Experiments of natural transition show a

wide range of transitionlocations from a high value of Re = 5.0 106 measured by

[18] to classical predictive theories,such as the e9 rule, which predict a value of2.0

106 [17]. This range of results is commonlyattributed to different noise levels in the

various experiments. We show later that the model canactually predict the entire

range of natural transition experiments by varying the model initialconditions. We


believe that the ability of this model to predict natural transition is a

uniquecharacteristic that is not demonstrated by other RANS models.

The other significant departure of the model predictions from the experiments

is in theovershoot of the friction coefficient that occurs at the very end of the

transition process andin the transition length. Predicting this overshoot and length

correctly is not as important aspredicting the transition location itself. Interestingly,

the model does show the characteristicovershoot behaviour found in experiments, but

over predicts the extent of this overshoot. Ourcurrent hypothesis is that this

discrepancy is largely caused by the use of boundary layerequations to solve the

velocity and turbulence variables, and is not due to the model itself. Theboundary

layer equations are based on the premise that streamwise second derivatives are

small.During the latter part of transition, as the flow suddenly becomes turbulent, the

boundary layer

Figure 2.

Comparison of predictions of transition in zero pressure gradientboundary layer for

T3A case.the symbols represent experimental data and thecurves are the predictions.
grows quite dramatically, and the assumption of small streamwise second derivatives

is not veryaccurate. Including these streamwise derivatives is expected to add

considerable diffusion andto significantly reduce the current excessive overshoot

behaviour. The fact that the overshooterror increases as the turbulence intensity

decreases is consistent with this behaviour, becausethe lower intensity transition

events are also more abrupt and therefore violate the boundarylayer approximations

the most severely.

Two other numerical predictions, both performed by [15], but using their

intermittencymodel and another model[6] respectively, are plotted in figure-2

for the T3A case. The model predictions are not particularly good for this

particular case. In contrast, the model ofSuzen and Huang captured the late stage of

the transition region (including the overshoot) verywell, but the transition onset was

delayed. The present calculation, as shown in figure-2, showsa smoother transition

behaviour. In other words, the present model predicts not only a correctposition of

transition onset, but also a fairly accurate length of the transition region.

The comparison of the Reynolds number based on momentum thickness, Re ,

and shapefactor, H, with experiments and results from other numerical models are

shown in figures-3and 4 respectively for the T3A case. It can be noted that the present

prediction always remainsbetween the other two simulations. The fact that the model

of Suzen and Huang has a slightlater prediction of the onset of transition is also

visible from the H curves.

The streamwise mean velocity profiles calculated using the present model are

compared atfour downstream stations with the experimental T3A data and the

predictions of Suzen andHuang in figures-5(a)(d). It is evident from these figures


that the model predictions agree very well with the experimental data even for the

detailed mean velocity profiles.

4.2. Flat plate boundary layer in non-zero pressure gradient:

Model predictions of transition in two variable pressure gradient boundary

layers are compared with the experimental data of [3] in figure-6. The initial

turbulence levels of these two cases are 3.0% (T3C3) and 6.6% (T3C1) respectively.

For both these flows, the pressure gradientis initially negative (favourable) and then

positive (adverse) in a profile that was designed to roughly approximate the flow over

a turbine blade. The pressure gradient profiles for the two

Figure 3. Comparison of Re for T3A case.


Figure 4.Comparison of H for T3A case.

different cases are roughly equivalent. The initial conditions are listed in table 2 and

are fully determined by the initial experimental data. It is very difficult to use

correlations to predict this variable pressure gradient flow because the pressure

gradient does not correspond closely to any single Falkner-Skan situation. In addition,

the RANS models tested by [13] showed some difficulty with these cases.

The model predicts the transition location in these complex variable pressure

gradientboundary layers quite well. Note that in both cases the initially favourable

pressure gradient has delayed transition compared to the flat plate case. The pressure

gradient data are not provided beyond the range of the experimental data, so we

cannot continue the 3% case beyond the experimental data. Predictions given by [15]

are compared with the present calculation for the T3C1 case again in figure-7.Figure-

8 shows the comparison of Re computed by the present model and the two other
Figure 5.

Mean streamwise velocity profiles for T3A case:

(a) Re = 1.35 105 ;(b) Re = 2.04 105 ; (c) Re = 2.74 105 ; (d)

Re = 4.19 105 .
Table 2.

Initial flow parameters at the leading edge of a variable pressure gradientflat

plate boundary layer.

( ) (%) Source

5.9 160 6.6 ERCOFTAC, T3C1

3.7 100 3.0 ERCOFTAC, T3C3

models with the experimental data for the T3C1 case. Predictions of the shape factor

profiles using the three models are plotted in figure 9 along with the experimental

data. Again the present model exhibits excellent performance in calculating the shape

factor in the transition region.

4.3. ACOUSTIC EFFECTS ON NATURAL TRANSITION:

It was mentioned previously that acoustic effects (wind tunnel noise) can have

a fairly strong influence on natural transition. The experiment of [18] is commonly

assumed to be largely free of acoustic noise, and therefore gives a high quiet

transition location of Rex,tr = 5.0 106 .Our


Figure 6.
Transition in non-zero pressure gradient boundary layer at variousinitial turbulence
intensities. The symbols represent experimental data and the curves are the present
results.

Figure 7.

Comparison of predictions of transition in non-zero pressure gradientboundary layer


for T3C1 case. The symbols represent experimental data and thecurves are the
predictions.
initial prediction of Re, = 2.0 106 is close to the low (noisy) end of the

experimental range.The experiments of [14] attempted to reduce noise levels, but

were clearly not as successful inthis respect as the later experiments of Well.

Interestingly, the turbulent scalar potential represents irrotational fluctuations

(see[9]),and could be a very good measure of the average acoustic forcing. Our initial

natural transition prediction (figure-1) was with relatively high noise levels of =

2/3, and is actually quite close to noisy experiments. If one reduces the initial

scalar potential to a value of 0.12, the on set of the transition will be delayed to

5.0 106 ,which agrees with[18].One can also match the

Figure 8.Comparison of Re for T3C1 case.


Figure 9.Comparison ofHfor T3C1 case.

data of [14] Re, = 2.8 106 in table 1 by using = 0.4. All three simulations

are shown inthe C Re curve plotted in figure-10. This demonstrates the ability of

the model to accurately capture the effects of noise on natural transition. The effects

of noise on bypass transition arepresent but are less significant.

4.4. Relaminarization in turbulent channel flow :

Almost as critical as transition is the process of relaminarization, where a

turbulent flow decays to a laminar state. Relaminarization can occur because of strong

stratification, strong rotation,or in the case of channel flow, due to a reduction in the

driving pressure gradient. Figure-11 shows predictions of the skin friction in channel

h
flow for various bulk Reynolds numbers Re = u ,where h is the channel half height

and is the average velocity. The figure also has dashed curves
Figure 10.
The acoustic effects on the natural transition at Tu= 0.03%. The dashed curves are
fully laminar and fully turbulent correlations.

Figure 11.

Predictions of the skin friction in channel flow for various bulk Reynolds numbers.
The dashed curves are fully laminar and fully turbulent correlations.
8
for the laminar exact result C = 3 Re1 and the turbulent correlation of [4]

C = 0.044 Re0.227.Here the skin friction coefficient is calculated using

1
C = (w)/ (2 u20 ), where wis the wall shear stress and u0 is the centre line

velocity of the channel. The model shows a clear relaminarization at

Re = 1700(Re = 120). Experiments and DNS [7] indicate a

transitionReynolds number of close to 1000 (Re = 60). So the model could be

improved. However,a theoretical study by [8] has shown that the linear instability for

a channel flow occurs atRe = 1600(Re = 110). It is our observation that this

relaminarization location is strongly influenced by the low Reynolds number

behaviour of the model (particularly the dissipation and pressurestrain term), and not

the transition term itself.


5. CONCLUSION

The ability of the turbulent potential model to predict transition in a wide

variety of boundarylayer flows has been demonstrated. This includes the ability to

predict everything from naturaltransition (a first for RANS-based models) to large

free-stream turbulence intensities. The modelis also demonstrated to be able to

accurately calculate the transitional flat plate boundary layerunder favourable and

adverse pressure gradients. The ability of this model to reflect the effect ofnoise on

the natural transition is also shown by comparison with corresponding experimental

dataunder different conditions. Finally, the relaminarization in turbulent channel flow

is successfullycaptured using this model. The computation cost of the proposed

method is comparable tothe widely used two-equation models such as /. However,

unlike /, the model does notassume equilibrium between the developing instability

and/or turbulence fluctuations, and it isthis non-equilibrium nature to which we

attribute these successful predictions.


CHAPTER 4
SCALING OF LYAPUNOV EXPONENTS IN HOMOGENEOUS
ISOTROPIC TURBULENCE

INTRODUCTION

One of the defining characteristics of turbulence is that it is unstable, with

small perturbations to the velocity growing rapidly. Indeed, turbulent flows in closed

domains appear to be chaotic dynamical systems [11]. The result is that the evolution

of the detailed turbulent fluctuations can only be predicted for a finite time into the

future, due to the exponential growth of errors. In a chaotic system, this prediction

horizon is inversely proportional to the largest Lyapunov exponent of the system,

which is the average exponential growth rate of typical linear perturbations. The

maximum Lyapunov exponent is commonly used to characterize the chaotic nature

of a dynamical system [7]. In a turbulent flow, the maximum Lyapunov exponent is

thus a measure of the strength of the instabilities that underlie the turbulence, and its

inverse defines the time scale over which the turbulence fluctuations can be

meaningfully predicted.

Lyapunov exponents in chaotic fluid flows have been estimated

experimentally since the work of Swinney [23], using indirect methods. In numerical

simulations, however, Lyapunov exponents can be determined directly by computing

the evolution of linear perturbations. This has been done for weakly turbulent Taylor

Couette flow [21] and very low Reynolds number planar Poiseuille flow [11].

Remarkably, to the authors knowledge, Lyapunov exponents have not been

determined for isotropic turbulence, a shortcoming corrected in this paper.


Homogeneous isotropic turbulence is an idealized turbulent flow that has been

extensively studied both experimentally [3, 4, 14] and using numerical simulations [2,

10, 19, 22]. It is valuable as a model for the small scales of high Reynolds number

turbulence away from walls [6]. It has been speculated that in isotropic turbulence, the

maximum Lyapunov exponent scales with the inverse Kolmogorov time scale [5],

suggesting that the dominant instabilities occur at Kolmogorov length scales as well.

If true, then a study of the maximum Lyapunov exponent and the associated

instabilities in homogeneous isotropic turbulence will be applicable to a wide range of

flows.

This paper focuses on how the maximum Lyapunov exponent and hence the

predictability time horizon scale with Reynolds number and computational domain

size of a numerically simulated homogeneous isotropic turbulence. The speculation

that should scale as the inverse Kolmogorov time scale [5] is in agreement with

an estimate from a shell model [1]. However, this scaling has not been directly tested

in direct numerical simulations.

In addition, in the process of computing the maximum Lyapunov exponent in

a direct numerical simulation, one necessarily computes the linear disturbance that is

most unstable (on average). This can be used in the short-time Lyapunov exponent

analysis, as introduced in [21], to characterize the nature of the instabilities. This will

be pursued here for isotropic turbulence. The remainder of this paper includes a

brief review of Lyapunov exponents and how they are computed in numerical

simulations (section II) followed by a description of the direct numerical simulations

studied here (section III). The results of a scaling study of the Lyapunov exponents

are given in section IV, and a short-time Lyapunov exponent analysis is presented in

section V, followed by concluding remarks in section VI.


4.1 LYAPUNOV EXPONENT ANALYSIS

Two important characteristics of chaotic dynamical systems for the purposes

of the current study are that 1) solutions evolve toward a stable attractor, and 2)

solution trajectories on the attractor are unstable so that near-by trajectories diverge

exponentially. The rate of this exponential divergence is characterized by the

Lyapunov exponents, whose characteristics are recalled briefly here. Further details

can be found in [21]. In addition, the use of Lyapunov exponents in the analysis

presented in the paper is described.

A. Evolution of Linear Perturbations

Consider a solution trajectory () of a chaotic system. The solution will

evolve toward an attracting set in phase space (the attractor); in turbulence this

corresponds to the solution evolving to a statistically stationary state. Let (0 ) at

some arbitrary starting time t0 be on the attractor, and consider an infinitesimal

perturbation u(t0) of the solution at time t0, and its evolution in time. The Lyapunov

exponents describe the growth or decay of the magnitude of u. In particular, the

multiplicative ergodic theorem [16] implies that the limit

1 ()
= lim ( ) (1)
(0 )

exists and is called a Lyapunov exponent. There is a spectrum of possible Lyapunov

exponents, depending on the solution u(t0) and the perturbation u(t0) at the starting

time. However, for almost all u(t0), = 1the largest Lyapunov exponent, and, due to

round-o error and other sources of noise, in practical computations, = 1 for all

u(t0). Furthermore, the Lyapunov spectrum (1> 2> 2> ) does not depend on
u(t0); it is instead a property of the dynamical system. See the review by Eckmann et

al. [7] for an introduction to the theory.

In addition, in practical computations as discussed above, we expect that in the

limit t

() 1 ()

(()) (()) (2)
() ()

and depend only on the solution at t, and not on the starting conditions
where

u(t0) and u(t0). The perturbation


is the disturbance that grows most rapidly in the

long run, growing at the average exponential rate . It is defined by the fact that

itslong-time average growth rate forward in time is and when the evolution is

backward in time the long-time average growth rate is . The short-time Lyapunov

exponent is simply the instantaneous exponential growth rate of


.

Because and
depend only on the solution at the current time, they can be

used as diagnostics for the instabilities responsible for a system being chaotic. In

particular, when is large, the underlying system is particularly unstable, and at that

time the Lyapunov disturbance


is rapidly growing. Thus by seeking out times

when is large, and by analyzing the solution u and the Lyapunov disturbance
at

that time, we can characterize the important instabilities. This is the short-time

Lyapunov exponent analysis described by Vastano& Moser [ 21].

In this paper we will be concerned with the scaling of with Reynolds number

and with the chaotic instabilities revealed by short-time Lyapunov exponent analysis.
4.2 SIMULATIONS

To simulate the base field, we solve the three-dimensional incompressible

Navier-Stokes equations on a cube of dimension L=2, with periodic boundary

conditions, to obtain a computational approximation of homogeneous isotropic

turbulence. Turbulence is maintained by introducing a forcing term to the Navier-

Stokes equations which only acts at large scales. The forcing formulation is described

in section III A. The Navier-Stokes equations are solved using a Fourier-Galerkin

spatial discretization with N modes in each direction, and the vorticity formulation of

Kim et al. [12]. This formulation has the advantage of exactly satisfying the

continuity constraint while eliminating the pressure term. A low-storage explicit third-

order Runge-Kutta scheme [20] is used for time evolution.

To compute the Lyapunov exponents, we compute the growth rate of a linear

perturbation added to the base field. This perturbation satisfies the linearized Navier-

Stokes equations:

1 2
+ ( + ) = + (3)

= 0, (4)

Where ui is the base field and ui is the disturbance field. The disturbance equations

are solved using the same numerical scheme as the Navier-Stokes equations. Note that

the forcing is applied only to the base field and not the perturbation. The

implementation of both the base and disturbance field solvers were verified using the

method of manufactured solutions.


A. Forcing

The goal of the forcing is to inject energy into the large-scale turbulence so

that the isotropic turbulence will be stationary. Forcing is applied to Fourier modes

with wave number magnitudes in a specified range, and is designed to produce a

specified rate of energy injection (forcing power), which, when the system is

stationary, will be the dissipation rate. By specifying the wave number range being

forced, forcing power and viscosity, the integral scale, turbulent kinetic energy and

Reynolds number can be controlled. The energy injection is accomplished by the

introduction of a forcing term fito the Navier Stokes equations:

1 2
+ = + + (5)

= 0 (6)

Following [20], in the Fourier spectral method used here, the Fourier transform of the

is specified in terms of the velocity Fourier transform as


forcing

() = ||2 ().
(7)

Case kfmin kfmax v N R /


1 0 2 0.0235 64 1.43 37.92 455.2
2 0 2 0.0113 96 1.58 58.34 123.8
3 0 2 0.0056 128 1.67 85.68 118.0
4 0 2 0.0038 192 1.70 106.33 51.2
5 0 2 0.0026 256 1.77 130.43 51.3
6 0 2 0.0010 512 1.82 211.76 69.5
7 2 4 0.0093 128 0.71 37.74 277.1
8 4 8 0.0037 256 0.35 37.31 72.1

TABLE I. Parameters defining the eight direct numerical simulations performed to

study Lyapunov exponent scaling. Values of are quoted in units in which the

domain size is 2, and averaging times are normalized by eddy turnover time.
Given that ui is a Navier-Stokes solution, fi is guaranteed to be divergence-

free. The coefficient in the above is determined as a function of time so that the

forcing power is the target dissipation rate T. Since the forcing is applied only to a

range of wavenumbers, this yields

1

= { ( ||2 () ()) || (8)
||
0

where denotes the complex conjugate, and kfminand kfmax are the bounds on the range

of wavenumbers being forced. In the Fourier transform of the Navier-Stokes

equations, the viscous term has the same structure as fi, so this forcing can be

interpreted as a negative viscosity acting in the specified wavenumber range. The

combined forcing and viscous term is then( )||2 (). In the numerical

solution of the Navier-Stokes equations, this combined term is treated in the same

way as the viscous term would be. Note that fi is just a nonlinear function of ui, so

there is no externally imposed stochasticity.

B. Simulation Cases

To investigate the scaling of the maximum Lyapunov exponent with both

Reynolds number and the ratio of the computational domain size L to the integral

scale , eight simulations were performed. These are summarized in table I. To study

the scaling of with Reynolds number, six cases where simulated with the same

forcing wavenumber range and .This resulted in approximately the same integral

scale in each case. The Reynolds number was manipulated by changing the viscosity.

To study the potential variation of with domain size normalized by integral scale,

the domain size was kept fixed at 2 and the integral scale was changed by adjusting

the forced wavenumber range, while keeping the Reynolds number approximately
fixed. In all cases kmax> 1, where kmax is the maximum resolved wavenumber, and

is the Kolmogorov scale. In a refinement study, this was found to be sucient to

obtain resolution independent values of .

For each case, the simulations were run until the base solution became

statistically steady and then the statistics were gathered by time averaging over a

period Tavgas reported in tableI. The simulation was confirmed to be stationary by

verifying the convergence of the viscous dissipation to and the statistical

convergence rates of q2and .

4.3. SCALING OF LYAPUNOV EXPONENTS

Of primary concern here is the dependence of the maximum Lyapunov

exponent on the Reynolds number and on the domain size. To address this, the

maximum Lyapunov exponent , the integral scale () and Reynolds number based on

the Taylor micro-scale (Re) are needed, along with their uncertainties. Based on the

assumption of isotropy, the latter two were determined to be = 0.15 3 / and

= 2 5/(3) [17]. Thus the two statistical quantities that need to be computed

from the DNS are and q2. Both are determined as a time average over averaging

time Tavg (see table I), and the standard deviations of the uncertainty due to finite

averaging time were determined using the technique described by [15]. The values of

, q2 and their standard deviations are given in table II. The standard deviations of the

derived quantities L and Re are determined simply as = (0.225/)2 and

= 5/(3)2 , where for it is assumed that 2 / 2 1. Note that since

for each simulation and are specified, there is no uncertainty in their values.
Case

1 4.51 0.107 0.0922 0.0038

2 4.80 0.160 0.1075 0.0046

3 4.99 0.075 0.1177 0.0032

4 5.05 0.044 0.1231 0.0040

5 5.19 0.046 0.1304 0.0034

6 5.28 0.084 0.1599 0.0048

7 2.82 0.008 0.0941 0.0019

8 1.76 0.001 0.0945 0.0021

TABLE II.Values of q2 and the maximum Lyapunov exponent , along with the

standard deviation () of the sampling uncertainty. Values of q2 are quoted in units in

which the domain size is 2 and = 1, and is normalized by the Kolmogorov time

scale .

The dependence of the maximum Lyapunov exponent in Kolmogorov units on

Reynolds number is shown in figure 1, including uncertainties expressed as the

standard deviation. If the hypothesized scaling of the Lyapunov exponent on

Kolmogorov time scale were correct, these data would, within their uncertainty, fall

along a horizontal line. However, this does not appear to be the case. Indeed,

appears to be growing with Re. Also, shown in figure 1 is the dependence of scaled

Lyapunov exponent on domain size at constant Reynolds number. These data do

appear to be consistent with the hypothesis that the Lyapunov exponent does not

depend on the domain size.

To make these scaling observations quantitative, Bayesian inference is used to

infer the coecients and in a scaling relationships of the form



= 1 1 = 2 (/)2 , (9)

given the data and its uncertainties. These scaling relationships serve as the model

for the inference. In Bayesian inference for this problem, the joint probability

distribution (, |d) of the parameters and conditioned on data d (shown in table

I) is sought. Bayes theorem gives this conditional probability as:

(, |) (|, )(, ) (10)

FIG. 1. Dependence of the Lyapunov exponent scaled in Kolmogorov units ( ) on

(a) the Taylor scale Reynolds number Re and (b) the ratio of the integral scale to the

domain size /L, from the data in table II. The error bars on the data (in blue)

represent one standard deviation. Also shown are the outputs of the models (9) (in

red) calibrated with Bayesian inference, with the dark and light gray shading

representing variations of one and two standard deviations respectively.

(a) (b)

where (d|, ) is the likelihood and (, ) is the prior. The likelihood is the joint

probability density for the observed quantities evaluated for the observed values of

these quantities, as determined by the model with parameters and , and given the

uncertainties in the data. The prior represents our prior knowledge about the

parameters, independent of the data.


The data are statistical averages obtained from direct numerical simulations.

The primary source of uncertainty in such data is statistical sampling error. The

central limit theorem implies that in the limit of large samples, the uncertainty

associated with sampling error is normally distributed with zero mean. Therefore, to

formulate the likelihood, the data are assumed to have Gaussian uncertainty with

standard deviations as reported in table-II. The probability distribution for the ith

observation of the value of as predicted by the models is thus given by

2

1 ( 1 1 )
( |, , ) = exp [ ] (11)
2 22

wherexi is the independent variable (Rei or /L, depending on which scaling relation

is being inferred) of the ith observation and i is the standard deviation in associated

with the ith observation. However, there are also uncertainties in the values of the

independent variables x, as determined from the DNS, again with a Gaussian

distribution and standard deviation for the ith observation of xi. In this case, the

probability distribution of the independent variable x given the observation xi is

1 ( )2
(| ) = exp [ ] (12)
2 22

The conditional distribution of given the parameters and the observed independent

variable is then given by

( |, , ) = (|, , )(| ) (13)


Finally, to obtain the likelihood, (13) is evaluated at = i and the uncertainties in

each observation are assumed to be independent (an excellent assumption), yielding:


(|, ) = ( = |, , ) (14)

To inform the prior, we consider the range of time scales in the turbulence.

The largest is the eddy turn-over time, which is proportional to q2/, and the smallest

is the Kolmogorov time scale /.The ratio of the turnover to the Kolmogorov

times scales as Re. Therefore, the Lyapunov exponent scaling with the turn-over

time would imply = 1 and scaling with the Kolmogorov scale would imply =0.

However, theoretical arguments suggest that the Lyapunov exponent scales with the

Kolmogorov time scale [5] (=0), and we need to allow for the possibility that this

assessment may be in error in either direction. The bounds on the range of plausible

values of were therefore extended to 1 1, and a uniform distribution over this

range was used as a prior for . somewhat arbitrarily, the same range was used for the

prior in the domain size scaling relationship. The parameter is a positive definite

scaling parameter, and so following Jaynes [9], a Jeries distribution ()1/ is used

as an (improper) prior. Finally, the priors for and are independent so

(, ) = ()().

Given the likelihood and prior described above, and the data in Table II,

samples of the posterior distribution were obtained using a Markov-chain Monte

Carlo (MCMC) algorithm [8] as implemented in the QUESO library [13, 18]. The

resulting samples were used to characterize the joint posterior distribution of and

for both the Reynolds number and domain size scaling model, as shown in figure-

2and 3. Notice in these figures that the joint distribution has probability mass
FIG. 2. Posterior PDFs for 1 and 1 in the Reynolds number scaling model (9).
Shown are the marginal distributions of both parameters along with contours of their
joint distribution.

concentrated in thin diagonally oriented regions, showing that uncertainty in and

are highly correlated. Indeed, the uncertainty in the s is as large as it is because

changes in can be compensated for by changes in so that the model still fits the

data. The MCMC samples were also used to determine the uncertainty in the model

predictions for the Lyapunov exponent as a function of Reynolds number and domain

size, with the results plotted in figure 1, along with the data. From this, it is clear that

the scaling models as calibrated are consistent with the data and their uncertainty.

The marginal posterior distribution for in the Reynolds number scaling

relation shows that the most likely values of are between about 1/4 and 1/3, with the

possibility that the value is zero essentially precluded. This is remarkable since it

suggests instability time scales that will become increasingly faster than Kolmogorov

with increasing Reynolds number.


FIG.3. Posterior PDFs for 2 and 2 in the domain size scaling model (9). Shown are
the marginal distributions of both parameters along with contours of their joint
distribution.
The origin of this fast time scale is currently unclear. One possibility to

consider is that this fast instability time scale arises as an artifact of the time

discretization of the DNS. However the DNS time step in Kolmogorov units

t/ Re 1/2, so if the Lyapunov exponent were scaling with the DNS time step,

would be 1/2, which is also essentially precluded by the posterior distribution. The

time discretization thus appears to be an unlikely origin of the observed Reynolds

number scaling. This was also verified by running a time refinement study where

was found to be invariant to changing t.As with the time step, interest in the

computational domain size arises because of concern that computational artifacts not

impact our Lyapunov exponent analysis. The posterior distribution of in the domain

size scaling relationship (figure-3) shows that =0 is highly likely, with the most

probable values of ranging from 0.05to 0.05. If there is an eect of the domain

size, the data indicates that it is extremely weak. It therefore appears that the

Lyapunov exponent Reynolds number scaling discussed above and the short-time

Lyapunov exponent analysis presented in section-V are unaected by finite domain

size eects.
4.4. SHORT-TIME LYAPUNOV EXPONENT ANALYSIS
As discussed in section II, both the disturbance field (u) used to compute the

Lyapunov exponent and its instantaneous exponential growth rate () depend only on

the instantaneous Navier-Stokes velocity u, not on the initial disturbance. In short-

time Lyapunov exponent analysis, we study and u to learn about the instabilities

responsible for the chaotic nature of turbulence.

First, consider the time evolution of the exponential growth rate , which is

shown in figure-4 for Re = 37 and 210 (cases-1 and 6 respectively), normalized by .

Note that in both cases takes large excursions from the mean, of order 3 times the

mean value. However, the variations in occur on a much shorter time scale and the

large excursions seem to occur more often in the high Reynolds number case. The

time scale on which varies appears to decrease somewhat faster than the

Kolmogorov time scale with increasing Reynolds number, as when plotted against

t/, still varies faster for case-6 (figure-5). At the same Reynolds number (cases-1

and case-8), the variability of decreases sharply with increasing relative

computational domain size L/. The fact that the time scale of the instability, as

measured by the Lyapunov exponent, decreases faster than the Kolmogorov time

scale suggests that the instability processes are acting at spatial scales near the

Kolmogorov scale. In this case, a simulation with a larger domain size relative to

intrinsic turbulence length scales would include a larger sample of local unstable

turbulent flow features, resulting in smaller variability in . In comparing case-8 with

case-1, the relative volume increases by a factor 64, suggesting that the variability of

should be about a factor of 8 smaller in case-8 than in case-1, which is indeed

consistent with the data.


At the peaks in , the growth of the disturbance energy is particularly rapid,

and the question naturally arises as to what is special about these times. To investigate

this, the spatial distribution of the magnitude of the disturbance energy density is

visualized in figure 6 at three times, just before the beginning of a peak in , a time

half way up that peak and at the peak (tq/ = 9.58, 9.85 and 9.89 in figure 4). Notice

that before the rapid growth of into the peak, the energy in the disturbance field is

broadly distributed across the spatial domain. Half way up the peak, the distribution is

much more spotty, and finally at the peak, the disturbance energy is primarily focused

in a small region, appearing in the lower left corner of figure 6(c). The contour levels

in these images were chosen so that the contours enclose 60% of the disturbance

energy, implying that 60% of the disturbance energy is concentrated in the small

feature in the lower left of figure 6(c). Another indication of the dominance of the

disturbance feature in figure 6(c) is that the contour level needed to enclose 60% of

the energy is about 2500 times the mean disturbance energy density, while in figure

6(a) the contour is only about 15 times the mean. Clearly the growth of the

disturbance field in this concentrated area is responsible for the peak in . However,

the spatially local exponential growth rate of the disturbance energy |u|2 |u|2/t is

not particularly large there, large values of this quantity are distributed broadly across

the spatial domain. It seems, then, that the large peak in is due to a local disturbance

that is able to grow over an extended time until it dominates the disturbance energy,

so that the disturbance is localized in a region of relatively large growth rate. This is

presumably unusual because it requires that the local unstable flow structure

responsible for the disturbance growth persists for a long time.

It is of interest to investigate the turbulent flow structures responsible for the

large localized disturbance energy at the peak in . In the region where u is


localized, the base field exhibits a pair of co-rotating vortex tubes (figure-7). As

shown in figure-8, the disturbance vorticity is localized on the vortex tubes, with

regions of opposite signed disturbance vorticity to one side or the other of each vortex

tube. This disturbance, when added to the base field would have the effect of

displacing each vortex tube along the line between the positive and negative peaks in

the disturbance vorticity associated with each tube. The instability then appears to be

one associated with slowing (speeding up) the co-rotation of the vortex tubes while

they move away from (toward) each other. Note that the disturbance equations, being

linear and homogeneous, are invariant to a sign change, and so the sign of the vortex

displacement is indeterminate. Such an instability of co-rotating vortices is

reminiscent of the pairing instability in two-dimensional mixing layers.

(a) Case 1, Re = 37 (b) Case 8, Re = 37, large L/

(c) Case 6, Re = 210 (d) Case 6, Re = 210, (zoomed in)

FIG.4. Short-time Lyapunov exponent scaled by . In (d), the time axis is expanded to

zoom in on the peak indicated in (c), and symbols show the times at which the images

in figure 6 were obtained.


CONCLUSION

The results of the scaling study (section-IV) show definitively that, at least

over the Reynolds number range studied, the Lyapunov exponent does not scale like

the inverse Kolmogorov time scale, as had been previously suggested [1]. Instead,

increases with Reynolds number like Re for in the range from 1/4 to 1/3. Further

note that the analysis of Aurellet al. [1] indicated that a correction for the

intermittence of dissipation would yield < 0, also inconsistent with the current

results.

(a) Case 1, Re = 37 (b) Case 6, Re = 210

FIG. 5.Short-time Lyapunov exponent / with time scaled by .

If positive scaling holds to much higher Reynolds numbers, it would be

remarkable, as it would mean that there are instability processes that act on time

scales shorter that Kolmogorov. However, in the highest Reynolds number (Re=210)

simulation performed here, is still only 0.16. it is certainly possible that this

Reynolds number dependence of is a low Reynolds number effect, caused by

insufficient scale separation between the large scales and the scales at which the

instabilities act, and that the value will reach a plateau at some much higher Reynolds

number. Clearly, this scaling behavior of the maximum Lyapunov exponent is worthy
of further study. The current results suggest that the generally accepted and most

obvious scaling is not correct, and that, unfortunately, turbulent fluctuations are even

less predictable than previously thought.

The short-time analysis described in section V confirmed that the dominant

instabilities in turbulence act on the smallest eddies. Further, at Re=210, when the

instantaneous disturbance growth rate was the largest (about 3 times the mean), the

disturbance energy was highly localized, suggesting that it was a particular local

instability that was responsible for the rapid growth at that time. However, this was not

due to a particularly large local growth rate, as the logarithmic time derivative of the

spatially local disturbance energy was equally large in regions spread throughout the

domain. It may be that the localized instability we observed is not of particular

importance, except that the underlying structure in the turbulent field was especially

long-lived. None-the-less, studying it showed that one of the possible instability

mechanisms acting in turbulence is reminiscent of pairing instabilities of co-rotating

vortices, as in a mixing layer. In this, the short-time Lyapunov analysis pursued here

appears to be a valuable tool for the study of the instabilities underlying turbulence.

(a) tq/=9.58 (b) tq/=9.85 (c) tq/=9.89

FIG. 6 Contour of the magnitude of at three times leading upto the peak as

indicated in figure 4 for Re = 210 (case 6). The contour shown is that at which 60%

of the disturbance energy is enclosed by the contour. To achieve this, the contour

levels are (a) 15, (b) 25 and (c) 2500 times the mean disturbance energy density.
FIG. 7. Contour of the magnitude of the vorticity of the base field for Re = 210 (case

6) at the peak of in the region where the disturbance field is localized (box

highlighted in figure 6c). The contour level is 9.2 times the square root of the mean

enstrophy. The vortex tubes are co-rotating, with the direction of rotation indicated by

the black arrow.

FIG. 8.The magnitude of the vorticity (grayscale) and the disturbance vorticity

component normal to the plane (contour lines) in a plane perpendicular to and in the

middle of the vortex tubes shown in figure 7. For the disturbance vorticity, the red and

blue contours are of opposite signs.


CHAPTER 5
CONTROLLING BOUNDARY LAYER TRANSITION OVER A
SEPARATION BUBBLE: A COMPLEX-LAMELLAR APPROACH

INTRODUCTION

In this paper, we analyze the laminar to turbulent boundary layer transition

that occurs within a separation bubble along a flat plate. Boundary layer transition is

important, both economically and ecologically. Since drag increases when a boundary

layer transitions, and because an increase in drag must be overcome by an equal and

opposite production of thrust, flight vehicles and fluid powered devices consume more

fuel and generate greater amounts of greenhouse gas emissions because of these

boundary layer transitions [1]. Furthermore, because transitioning boundary layers are

prone to separate, they can dramatically affect the overall performance of high lift

devices and gas turbines [2, 3]. When a laminar boundary layer separates, transitions

into a turbulent flow and then reattaches downstream, the resulting phenomena is

called a separation bubble [4]. And these separation bubbles produce drag increases

that are greater than those that would otherwise be produced by transitioning but

attached boundary layers [5].

Early theoretical studies of boundary layer transition produced the stability


analyses of Rayleigh [6], Orr [7], Sommerfeld [8], Tollmien [9], Schlichting [10],
Taylor [11], and Lin [12], which identified where a steady laminar boundary layer
becomes linearly unstable. These analyses assumed that a flows stability depends
only on its local flow conditions at any given location. More recently, both Herbert
[13] and Dallmann [14] developed stability analyses that captured both the streamwise
development and upstream history of the resulting flow instabilities [15]. These works
have been validated through experimental studies [16, 17] and have helped to identify
how the stability of a laminar boundary layer influences transition location.
Experimental studies of separation bubbles exist, however many are limited by

the challenges of measuring separated flows [18]. Gaster [19] developed an

experimental criteria to determine the size of a separation bubble. Using this work and

others, Mayle [2] developed experimental correlations that would predict the sizes of

separation bubbles. More recently, Haggmark et al. [18] have performed a more

detailed analysis and visualization to determine overall bubble lengths.

Direct numerical simulations (DNS) have also been used to investigate

transition [4, 20, 21]. DNS, although computational expensive [22], have produced

some important studies of separation bubbles [4].Spalart and Strelets [23] simulated a

laminar boundary layer on a flat plate that detaches, transitions to turbulence,

reattaches and then evolves into a normal turbulent boundary layer. Alam and

Sandham [4] performed a simulation of a short separation and Balzer and Fasel [24]

simulated the effects that freestream turbulence has on the length of short separation

bubbles. Rist et al. [21, 25, 26] also studied how disturbances can be used to control

the size of separation bubbles for both short and long separation bubbles.

In this paper we will obtain the locations of separation, reattachment and the

start and end of transition for the flow over separation bubbles. We will verify our

belief that when the separation bubble reattaches, the turbulent boundary layer flow

will be fully developed. We will use a transition model derived from a complex-

lamellar decomposition of the velocity field across transition that exists between a

fully developed laminar and turbulent boundary layer flows. This transition model is

augmented with the experimental correlations of Mayle [2] and Narasimha [27].

These results will be compared with several different DNS simulations.


5.1. GOVERNING EQUATIONS

For a viscous, incompressible flow in the Cartesian coordinate system

(x, y, z) the conservation laws of continuity and linear momentum, and the vorticity

transport equation, can be written as

. = 0, (1)

1
= + 2 . (2)


= (. ) + 2 . (3)

where u D (u, v, w) are Cartesian velocity components, = ( , , )are

Cartesian vorticity components, is fluid density, p is fluid pressure, is kinematic

viscosity, and the material derivative is written as


= + (u. ) (4)

Complex-Lamellar Decomposition

The velocity field that satisfies the incompressible NavierStokes equations

can be decomposed into a complex-lamellar form in general [28], and the following

Monge transformation [29] in particular:

= + (5)

Here the velocity field is separated into two components, one potential and the

other rotational. The flexion vector [30]

2 = + (6)
is constructed from the potential , often called either Craigs circulation preserving

flexion-potential [31] or the HelmholtzRayleigh energy dissipation potential [32].

This energy dissipation potential also defines the viscous behaviour of the transport of

the velocity potential

.
= (7)
2

which defines the irrotational component of the velocity field.


= 1. (8)

is the transport of the classic drift function [3335]; and


= , (9)

is the transport of an enstrophy function E that is required for dimensional

consistency, and contains a source term that either creates or destroys enstrophy

along the particle paths. Given the closed curve c around a material surface s, with

line and surface elements land, the transport of E must satisfy the following

equation:


= 2 (. ), (10)

which, assuming all variables are single-valued, we can re-cast as


= 2 [ x]2 , (11)

which allows us to identify the source term as

= 2 [ x]2 , (12)

Augmented Laminar Flow

Any flow, laminar, turbulent or transitioning, can be decomposed into the

following complex-lamellar form:

= + + (13)

where a complex-lamellar velocity decomposition, equivalent to that which was

previously defined, Eq.(5), has been added to ulam, a baseline laminar flow.

However, here the enstrophy term is defined as

= , (14)

the difference between E, the enstrophy function present in the given flow of

interest, and that which would otherwise be found in a baseline laminar flow, Elam.

The transport of this enstrophy difference can be obtained from


= 2 [ x]2 (15)

which would require an appropriate application of initial and boundary conditions

to transform a baseline laminar flow into the particular flow of interest.

For a transitioning 2D boundary layer, the enstrophy difference, Eq. (14), is

interpreted as a change in enstrophy across transition, and constructed as


= 2 )
(2 , (16)

where2 is the square of the magnitude of the vorticity found in the transitioning

2
boundary layer, while is that which would otherwise be found in the baseline
laminar flow. Thus, at the end of the fully developed laminar boundary layer flow,

which occurs at x = xA, we have the boundary condition | = 0. While at x = xB,


the location where the boundary layer is assumed to be a fully developed turbulent

flow, the change in enstrophy becomes | = ( ) | .


Given that the definition of our complex-laminar decomposition, Eq. (13)

requires at the fully developed turbulent boundary layer location, xB,

[ x ] | = ( ) | , (17)

which is a change in vorticity at the end of transition. It can be shown [29], that for

a 2D boundary layer flow, the following drift function approximation can be made:


= 1, (18)


= 0, (19)


= [2 sin(1 ) + 3 (2 )]. (20)


in regions where v 0, 0 and u 0, or within a boundary layer but away from

the solid wall.

The approximate drift function, Eq. (18), when coupled with the change in

enstrophy, Eq. (16) allows us to rewrite Eq. (17) as

1 2 )
[ (2 ] | = ( ) | . (21)

At the ends of this region, we recognize the following boundary conditions:

0
2 2 when =
= {(2 2 ) , (22)
| when =

and assume the following series approximation:


2 2 2 2 )
( )
= ( | ( ). (23)
( )
=1

which is the change in enstrophy density generated as the flow transitions from

laminar to turbulent. To satisfy the boundary condition at x = xB, this approximation

requires that

= 1 (24)
=1

Finally, we can substitute the approximate enstrophy distribution, Eq. (23), into Eq.

(21), to produce the following relationship:



| = ( + ) | ( ), (25)
+ 1
=1

an equation that states that the velocity at the end of transition is a function of the

change in enstrophy between the fully developed laminar and turbulent flows, the

sum of the resulting turbulent and otherwise laminar vorticities, at the end of

transition, and the length of the fully developed laminar to fully developed turbulent

region itself, (xB xA). These results are dependent on the series coefficients Cn,

which in turn define the shape of the change in enstrophy density across transition.
5.2. MODEL OF INTERMITTENCY

From the end of the fully developed laminar boundary layer flow, xA, to the

beginning of the fully developed turbulent flow, xB, we can define the scalar factor

as


( )
= , (26)
( )
=1

which can be used to write the change in enstrophy density as

2 2 2 )
2 = ( | . (27)

It was shown [29] within the measurable extent of the intermittency region,

= = 0.75 = 0.25, the shape of the change in enstrophy density, is

equivalent to the shape of the universal intermittency distribution, . Where =

0.75 identifies the location where the mean flow is approximately 75 percent

turbulent and 25 percent laminar, and = 0.25 identifies its counterpart. The shape

of the universal intermittency distribution is defined as [27]

( )2
= 1 [0.412 ], (28)
2

wherext is the equivalent leading edge of the turbulent boundary layer, and the

location where the universal intermittency function becomes = 0. Therefore, we

can show that


( )
= = 0.25 = 0.75 (29)
( )
=1
Figure 1 illustrates how the extent of intermittency is small in length compared with

the overall region from xAto xB. As such, we chose to limit the series approximation

to three terms and generate three geometric equations using Mayles [2]

experimental correlation for the length of the constant pressure region,

= 4000.7

, (30)

along with the correlation to relate it to Narasimhas [27] extent of intermittency,

( )
= (31)
3.36

The following geometric equations are used for transition over a separation bubble:

3
( + 154.420.7

)
0.5 = , (32)
=1
( )

3
( + 99.4760.7

)
0.25 = , (33)
=1
( )

3 1
8.257 104 ( + 154.420.7

)
= , (34)
0.7
( )
=1

5.3. BOUNDARY CONDITION AT THE END OF TRANSITION

The fully developed turbulent boundary layer flow at xB is satisfied using

the Prandtl-von Karman power law velocity profile and the fully developed laminar

boundary layer flow at xA is satisfied using the Pohlhausen laminar velocity profile.

Substituting both the turbulent and laminar velocity profiles, into the

complex-lamellar relationship, Eq. (25), and with this relationship being evaluated

at the perpendicular location, y =x, we obtain


6 1
7 7
1 1 1 (2 + /6)
( ) ( 4 ) + 1
7 | 5 2
0.375 ( ( )) 5( )
[

3(2 /2)2
| |

3
(5)2 ( ) (5)3 ( ) 2

4(1 /6)3 3
|


+ 2 [ ( 1) ]
(5)4 ( ) +1
=1
]
1
7

|

=( 4/5
) (35)
0.375 ( ( ))

with a complex-lamellar ratio, R=xB/xA, and a laminar ration, RL = xt/xA.

In Fig. 1, we show a pictorial representation of the various spatial locations

that exist within our transition model as well as a sketch of the laminar an turbulent

boundary layer thickness that bound these regions.

Fig. 1. Transition model over a separation bubble bounded between xA and xB.
5.4. RESULTS

Our transition model, Eqs. (24) and (32) to (34) and section 4, is used to

determine the streamwise locations of separation, reattachment, and the start and

end of intermittency for the transitioning flow across several different separation

bubbles. We will present results that suggest that reattachment occurs when the

turbulent boundary layer becomes fully developed. The system of equations that

define our transition model is solved numerically using MATLAB (MATLAB

2010b, The MathWorks, Inc., Natick, MA) with a convergence tolerance of1012.

Multiple solutions are obtained for each test case, however only the solutions

assumed to be physically relevant, the ones consistent with the DNS simulations,

are presented.

Our first test case will compare the results obtained from our transition

model to those found in the DNS study by Cadieux et al. [36], which is based on the

work of Spalart and Strelets [23]. Since the freestream turbulence level imposed

upon these DNS simulations was assumed to be around Tu = 0.03%, we used this

turbulence level in the Abu-Ghannam and Shaws [37] momentum thickness

correlation,

( )
= 163 + [( ) ]. (36)
6.91

where

( ) = 6.91 + 12.75 + 63.64( )2 (37)

to determine a momentum thickness Reynolds number at separation. This

momentum thickness Reynolds number is used to identify the cross-stream height,

above the solid surface, at which we will apply our transition model. Thus, for a

freestream turbulence level of Tu = 0.03%, Abu-Ghannam and Shaws [37]


correlation, Eq. (36) produces a momentum thickness Reynolds number at

separation of / = / = 689. The fully developed laminar boundary layer

that is about to separate, at the location , can be modeled by the Pohlhausen

velocity profile [38] and the separation parameter = 12. Downstream, at ,

where the flow has reattached and becomes fully developed and turbulent, the

resulting boundary layer velocity profile is modeled by the Prandtl velocity profile

[39]. From Cadieux et al. [36] DNS data, we assume the Reynolds number

ateparation to be = 9.33 104 . Upon solving the system of equations that

form

Fig. 2. Change in enstrophy density with associated stream wise locations.

our transition model, we obtain the following solution:

C1 = 4.0995 C2 = 9.4692 C3 = 4.3697


RL = 1.3279 R = 1.6231 (38)

which predicts the effective leading edge of the turbulent boundary layer to be

located at = 1.23 x 105 and the fully developed turbulent boundary layer

flow to be established at = 1.51 x 105 .


In Fig. 2, we plot the change in enstrophy density, Eq. (26), between the end

of the fully developed laminar boundary layer flow, to the beginning of the

fully developed turbulent boundary flow, .

This figure allows us to identify a number of different streamwise locations

and interpret how the flow changes from a fully developed laminar to a fully

developed turbulent boundary layer flow. At , the change in enstrophy density

is zero and the laminar boundary layer flow is fully developed and assumed to

separate. Between and 0 , the change in enstrophy density grows

negatively, which implies that the boundary layer is now adjusting to the upstream

presence of the separation bubble. Thus while the boundary layer is assumed to

remain laminar, vorticity is being redistributed and the resulting flow is no longer

fully developed.Between 0 , the start of intermittency, and , the end of

intermittency, the shape of the change in enstrophy density is assumed to follow the

shape of the universal intermittency function, Eq. (29), which grows from zero to

one. In this region, the boundary layer flow is assumed to be changing,

intermittently, from laminar to turbulent until finally, at , the boundary layer is

fully turbulent. However, between and the flow is assumed to be

evolving until finally, at , the turbulent boundary layer flow becomes fully

developed. For the separation bubble modelled by this test case, Fig. 2 indicates that

the end of intermittency occurs at the same streamwise location as the start of the

fully developed turbulent boundary layer flow, or more simply, . This

implies that within this separation bubble the turbulent boundary layer flow

becomes fully developed at the end of transition. Furthermore, the result, that

, suggests that the turbulent boundary layer is fully developed at


reattachment. To verify this observation, we compared the location of with the

reattachment location, , identified from Cadieux et al. [36] DNS simulation.

This comparison is shown in Table 1.

Table 1.Comparison of the streamwise location of the start of the fully developed

turbulent boundary layer flow with Cadieux et al. [36] reattachment location.

Fully Developed Turbulent Cadieux et al. [36]


% Difference
Boundary Layer Flow Reattachment location

= 1.51 x105 = 1.53 x105 1.3

Table 2. Comparison of the separation bubble length between the fully developed

laminar to fully developed turbulent boundary layer flows with Cadieux et al. [36]

distance between separation and reattachment.

Distance from Fully Developed Cadieux et al. [36] Distance


Laminar to Turbulent Boundary from Separation to % Difference
Layer Flows Reattachment

= 5.77 x104 = 5.97 x104 3.4

The 1.3% difference between and supports the belief that the turbulent

boundary layer is fully developed when it reattaches.

To further validate our calculations, we compare the distance between our

fully developed laminar and turbulent boundary layer flows with the distance

between separation and reattachment from Cadieuxet al. [36] DNS simulation. As

shown in Table 2, the 3.4% difference between these distances implies that the

distance between the fully developed laminar and turbulent boundary layer flows,

i.e. , measures the length of the separation bubble.


Furthermore, in his experimental studies to characterize short and long

separation bubbles, Gaster [19] concluded that for short separation bubbles the

streamwise location of the end of transition occurred near the bubbless

reattachment location. Thus, we can conclude that the separation bubble identified

by our calculations is in fact a short bubble.

We confirm, with a second test case of higher freestream turbulence levels,

that the start of the fully developed turbulent boundary layer flow occurs at

reattachment. We compare the results obtained from our transition model to those

found in the Alam and Sandham [4] DNS simulation. From their DNS data, we will

impose a freestream turbulence level of Tu = 0:46% and a corresponding

momentum thickness at separation of = 315.

Again, we assume that the fully developed laminar boundary layer that is

about to separate, at the location , can be modeled by the Pohlhausen velocity

profile [38] and the separation parameter D 12. And again, downstream at ,

where the flow has reattached and becomes fully developed and turbulent, the

resulting boundary layer velocity profile is modeled by the Prandtl velocity profile

[39]. From Alam and Sandham [4] DNS data, we assume the Reynolds number at

separation to be = 1.54 x104 . Upon solving our transition model, we obtain

the following solution:

C1 = 4.1082 C2 = 9.4812 C3 = 4.3730

RL = 2.1517 R = 3.1858 (39)

which predicts the effective leading edge of the turbulent boundary layer to be

located at = 3.31 x104 and the start of the fully developed turbulent boundary

layer flow to be initiated at = 4.91 x104


Table 3.Comparing the streamwise location of the start of the fully developed

turbulent boundary layer flow with the reattachment location from Alam and

Sandham [4] DNS simulation.

Fully Developed Turbulent Cadieux et al. [4]


% Difference
Boundary Layer Flow Reattchament location

= 4.91 x104 = 4.85 x104 1.2

Table 4. Comparison of the separation bubble length between the fully developed

laminar to fully developed turbulent boundary layer flows with Alam and Sandham

[4] distance between separation and reattachment.

Distance from Fully Developed Alam and Sandham [4]


Laminar to Turbulent Boundary Distance from Separation to % Difference
Layer Flows Reattchament

= 3.37 x104 = 3.31 x104 1.8

In Table 3, we compare the start of the fully developed turbulent boundary layer

flow, , with the reattachment location, obtained from Alam and

Sandhams [4] DNS simulation.

For this test case, where the turbulence level was specified as Tu = 0.46%,

we have a 1.2% difference between the start of the fully developed turbulent

boundary layer flow, , and the reattachment location, . This result further

supports the belief that the start of the fully developed turbulent boundary layer

flow occurs at reattachment.

66
In Table 4, we again compare the distance between the fully developed

laminar and turbulent boundary layer flows with the distance between separation

and reattachment from Alam and Sandham [4] DNS simulation. The 1.8%

difference between these distances again supports the belief that the distance

between the fully developed laminar and turbulent boundary layer flows,

measures the length of the separation bubble.

In Fig. 3, we show that the change in enstrophy density, Eq. (26), grows in a

similar way to the previous test case, specifically in terms of how the flow develops

from the end of the fully develop laminar boundary layer flow, at , to the start

of the fully developed turbulent boundary layer flow, at . Again, the end of

intermittency, occurs at the same streamwise location as the start of the fully

developed turbulent boundary layer flow, , which implies that the separation

bubble calculated for this test case is also a short bubble.

In the next test case, we use our transition model to solve for three different

freestream turbulence levels, Tu=0.05, 0.5 and 2.5% and use the change in

enstrophy density, Eq. (26), to demonstrate how the freestream turbulence level

effects the reattachment location and length of the separation bubbles. We com-pare

our results to those found in Balzer and Fasel [24] DNS simulation.

For each freestream turbulence level, we again assume that at , the fully

developed laminar velocity profile is about to separate and can be modeled by

imposing the separation parameter = 12 within the Pohlhausen velocity profile

[38]. As well, downstream at , where the flow has reattached and becomes

fully developed and turbulent, the resulting boundary layer velocity profile is

modeled with the Prandtl velocity profile [39].

67
The first separation bubble is obtained from solving our transition model

with a freestream turbulence level of Tu = 0.05% which, using Abu-Ghannam and

Shaws [37] correlation, Eq. (36) results in a momentum thickness Reynolds

number at separation of = = 679. As well, from Balzer and Fasel [24]


| |

DNS data, we assume the Reynolds number at separation to be

= 1.173 x105 . Then when this configuration is solved, we obtain the

following solution:

Fig. 3. Change in enstrophy density with associated streamwise locations.

C1 = 4.0924 C2 = 9.4595 C3 = 4.3570

RL = 1.2578 R = 1.4905 (40)

which predicts the effective leading edge of the turbulent boundary layer to be

located at = 1.75 x 105 and the fully developed turbulent boundary layer flow

to be established at = 1.75 x 105 .

68
The second bubble is obtained with a freestream turbulence level of

Tu = 0.5%, a momentum thickness Reynolds number at separation, obtained from

Abu-Ghannam and Shaws [37] correlation, Eq. (36), of = = 506, and


| |

a Reynolds number at separation, assumed from Balzer and Fasel [24] DNS data, of

= 1.177 x 105 . Solving our transition model with this configuration, we

obtain the following solution:

C1 = 4.0742 C2 = 9.4343 C3 = 4.3601

RL = 1.2077 R = 1.3964 (41)

which predicts the effective leading edge of the turbulent boundary layer to be

located at = 1.42 x 105 and the start of the fully developed turbulent boundary

layer flow to be initiated at = 1.64 x 105 .

Finally, the third bubble has a freestream turbulence level of Tu = 2.5%.

From Abu-Ghannam and Shaws [37] correlation, Eq. (36), we obtain a momentum

thickness Reynolds number at separation of = = 218, and using the


| |

DNS data from Balzer and Fasel [24], we assume a Reynolds number at separation

of = 1.184 x 105 . When this configuration is solved within our transition

model, we obtain the following solution:

C1 = 3.9876 C2 = 9.3157 C3 = 4.3281

RL = 1.1118 R = 1.2162 (42)

69
Fig.4. Change in enstrophy density with associated streamwise locations illustrating
the changes in bubble length with changes in freestream turbulence level.
which predicts the effective leading edge of the turbulent boundary layer to be

located at = 1.32 x 105 and the fully developed turbulent boundary layer flow

to be established at = 1.44 x 105 .

We can use the distance from the end of the fully developed laminar to the

fully developed turbulent boundary layer flows, , to measure the length

of the separation bubble. Thus in Fig. 4, the change in enstrophy density, Eq. (26),

obtained for each freestream turbulence level is plotted. When the freestream

turbulence level is the lowest, at Tu = 0.05%, Fig. 4 shows that a separation bubble

with the longest length of = 5.77 x 104 is obtained. When the

freestream turbulence level is the highest, at Tu = 2.5%, the resulting bubble is the

shortest with a length of = 2.56 x 104 . This implies that by increasing

the freestream turbulence level, the length of the separation bubble will become

shorter. This trend is consistent with the DNS study of Balzer and Fasel [24].

70
In Table 5, when we compare the start of the fully developed turbulent

boundary layer flow, , with the reattachment location, obtained from

Balzer and Fasel [24] DNS simulation, the larger differences between these

locations, compared to previous test cases, is apparent. In Table 6, the distances

between the end of the fully developed laminar and the start of the fully developed

turbulent boundary layer flows is compared to the distance between separation and

reattachment from Balzer and Fasel [24] DNS simulation and again greater

differences, compared to previous cases, is also apparent. Although these larger

differences are not ideal, they are within our expectations, considering the

approximations that we imposed within our transition model. Namely, that

gradients in the cross-stream direction are assumed to be small, which may be

unlikely near reattachment.

Table 5. Comparing the streamwise location of the start of the fully developed

turbulent boundary layer flow with the reattachment location from Balzer and Fasel

[24] DNS simulation for three freestream turbulence levels of Tu = 0.05, 0.5 and

2.5%.

Fully Development Balzer and Fasel


Freestream
Turbulent Boundary [24] Reattachment % Difference
Turbulence Level
Layer Flow Location
0.05% = 1.75 x105 = 1.60 x105 9.4
0.5% = 1.64 x105 = 1.54 x105 6.5
2.5% = 1.44 x105 = 1.49 x105 3.4

Table 6. Comparison of the separation bubble length between the end of the fully

developed laminar to the start of the fully developed turbulent boundary layer flows

with Balzer and Fasel [24] distance from separation to reattachment for the different

freestream turbulence levels.

71
Distance from Fully Balzer and Fasel [24]
Freestream
Developed Laminar to Distance from separation %
Turbulence
Turbulent Boundary Layer to Reattachment Difference
Level
Flow
0.05% = 5.77 x104 = 4.27 x104 35
0.5% = 4.63 x104 = 3.63 x104 26
2.5% = 2.56 x104 = 3.06 x104 16

In the final test case, we present how changes to a disturbance amplitude can

produce a long separation bubble and a short bubble. We compare the results we

obtain using our transition model with those obtained in the DNS study by Rist

[25].

We can calculate the freestream turbulence level, Tu, that is associated with

the disturbance amplitude using

2 2
1/2( + )
= . (43)

From Rists [25] DNS study, we obtain a disturbance amplitude of = 106 and

105.2to calculate a freestream turbulence level of Tu = 0.00045% and a

disturbance amplitude of = 104 and 103.25 to calculate a freestream

turbulence level of Tu = 0.04%. For each freestream turbulence level, we can again

assume within our transition model that at , the fully developed laminar

boundary layer flow can be modeled by Pohlhausen velocity profile [38] and the

separation parameter = 12 and at , the fully developed turbulent boundary

layer flow can be modeled by the Prandtl velocity profile [39].

For the case with a freestream turbulence level of Tu = 0.00045%, Abu-

Ghannam and Shaws [37] correlation, Eq. (36) results in the momentum thickness

72
Reynolds number at separation of = = 703. From Rists [25] DNS
| |

data, we assume the Reynolds number at separation to be

= 1.24 x106 . Upon solving the system of equations associated with our

transition model, we obtain the following solution:

C1 = 16.9526 C2 = 48.0084 C3 = 30.0558

RL = 1.0742 R = 1.1477 (44)

which predicts the effective leading edge of the turbulent boundary layer to be

located at = 1.33 x106 and the fully developed turbulent boundary layer flow

to be established at = 1.42 x106 .

Fig. 5.Change in enstrophy density for long bubble produced with a freestream

turbulence level of Tu = 0.00045% and short bubble produced with a freestream

turbulence level of Tu = 0.04%.

For the larger freestream turbulence level of Tu = 0.04%, the momentum

thickness Reynolds number at separation, obtained from Abu-Ghannam and Shaws

[37] correlation, Eq. (36) is = = 684. And the Reynolds number at


| |

73
separation, from Rists [25] DNS data, is assumed to be = 1.29 x106 . Upon

solving this configuration, we obtain the following solution:

C1 = 3.8913 C2 = 9.1857 C3 = 4.2944

RL = 1.0221 R = 1.0435 (45)

which predicts the effective leading edge of the turbulent boundary layer to be

located at = 1.319 x106 and the start of the fully developed turbulent

boundary layer flow is initiated at = 1.35 x106 .

In Fig. 5, we plot the change in enstrophy density, Eq. (26), for both

freestream turbulence levels. For a freestream turbulence level of Tu = 0.04%,

shown as the dashed curve, the change in enstrophy density has similar growth

characteristics to the previous cases and in particular, the end of intermittency

location, , corresponding to the start of the fully developed turbulent boundary

layer flow, . This is not the case for the lower freestream turbulence level of

Tu = 0.00045%, represented as the solid line in Fig. 5. For Tu = 0.00045%, the

change in enstrophy density continues to evolve between and . This

implies that for this separation bubble the turbulent boundary layer flow at the end

of intermittency is not fully developed and only becomes fully developed further

downstream.

In his experimental studies, Gaster [19] concluded that for long separation

bubbles, reattachment occurred at a streamwise location further downstream then

the end of transition. Thus, we can verify that the turbulent boundary layer flow

becomes fully developed downstream of the end of intermittency by comparing

Table 7. Comparing the streamwise location of the start of the fully developed

turbulent boundary layer flow with the reattachment location obtained from Rists

74
[25] DNS simulation for freestream turbulence levels of Tu = 0.00045% and

Tu = 0.04%.

Fully Developed Rist [25]


Freestream
Turbulent Boundary Reattachment % Difference
Turbulence Level
Layer Flow Location
0.00045% = 1.42 x106 = 1.40 x106 1.4
0.04% = 1.35 x106 = 1.36 x106 2.2

Table 8. Comparison of the separation bubble length between the end of the fully

developed laminar to start of the fully developed turbulent boundary layer flows

with Rists [25] distance from separation and reattachment for the different

turbulence levels.

Distance from Fully


Freestream Rist [25] Distance from
Developed Laminar to %
Turbulence Distance from separation
Turbulent Boundary Layer Difference
Level to Reattachment
Flow
0.00045% = 1.8 x105 = 1.6 x105 13
0.04% = 6 x104 = 7 x104 14

the location of with the reattachment location, identified from Rist [25]

DNS simulation. This comparison, along with the comparison for Tu = 0.04%, is

made in Table 7.

The 1.4 and 2.2% difference for the freestream turbulence levels of

Tu = 0.00045 and 0.04%, respectively, supports the belief that the boundary layer

flow is turbulent and fully developed at reattachment for both short and long

separation bubbles. We can also conclude, for the separation bubbles obtained in

this case, that a freestream turbulence level of Tu = 0.00045% produced a long

separation bubble and Tu = 0.04% produced a short separation bubble.

75
We can calculate the distance between the fully developed laminar and

turbulent boundary layer flows to determine the length of the long and short

separation bubbles. For the long bubble, we obtain a length of

= 1.8x105 and for the short bubble a length of

= 6 x104 . These lengths are comparable with the distance between

separation and reattachment from Rists [25] DNS simulation, as shown in Table 8.

76
CONCLUSION

Within this work, we developed a new complex-lamellar description of the

incompressible flow that exists as a boundary layer transitions from a fully

developed laminar to fully developed turbulent flow. This transition model was

used to obtain the locations of separation, reattachment and the starts and ends of

intermittency for several different separation bubbles. The fully developed laminar

boundary layer was modeled by a Pohlhausen velocity profile while its counterpart,

a fully developed turbulent boundary layer was modeled by a Prandtl velocity

profile. Between these two fully developed boundary layer flows, the transitioning

flow is assumed to separate, reattach and form a separation bubble. Transition over

several different separation bubbles, both long and short, were analyzed and

compared to several different DNS studies. In each of the cases analyzed, the

transition model predicted that the separation bubble reattaches when the turbulent

boundary layer flow becomes fully developed.

77
BIBLIO GRAPHY

[1] Abu-Ghannam B J and Shaw R 1980 Natural transition of boundary layers


the effects of turbulence, pressure gradient, and flow history J. Mech. Eng.
Sci. 22 21328

[2] Are S, Zhang X and Perot J B 2001 Accuracy and conservation properties of
a three-dimensional unstructured staggered mesh scheme for fluid dynamics
Int. J. Rotating Machinery at press

[3] Coupland J 1990a ERCOFTAC Special Interest Group on Laminar to


Turbulent Transition and Retransition T3A and T3B Test cases Coupland J
1990b ERCOFTAC Special Interest Group on Laminar to Turbulent
Transition and Retransition T3C Test cases

[4] Dean R B 1978 Reynolds number dependence of skin friction and other bulk
flow variables in two-dimensional rectangular duct flow J. Fluids Eng. 100
21523

[5] Dhawan S and Narasimha R 1958 Some properties of boundary layer flow
during the transition from laminar to turbulent motion J. Fluid Mech. 3 418-
36

[6] Launder B E and Sharma B I 1974 Application of the energy dissipation


model of turbulence to the calculation of flow near a spinning disc Lett. Heat
Mass Transfer 1 1318

[7] Keefe L, Moin P and Kim J 1992 The dimension of attractors underlying
periodic turbulent Poiseuille flow J. Fluid Mech. 242 129

[8] Orszag S 1971 Accurate solution of the OrrSommerfeld stability equation J.


Fluid Mech. 50 689703 [9] Perot J B 1999 Turbulence modeling using body
force potentials Phys. Fluids 11 264556

78
[10] Perot J B and Taupier J 2000 Modeling three-dimensional boundary layers
using the turbulent potential model AIAA 2000-0914 (Reno, Nevada,
January 2000)

[11] Perot J B and Wang H 1999 Modeling separation and reattachment using the
turbulent potential model Proc. 4th Int. Conf. on Turbulence Modeling and
Measurements (Corsica, France, May 1999)

[12] Reynolds W C and Kassinos S C 1995 One-point modeling of rapidly


deformed homogeneous turbulence Proc. R. Soc. A 451 87104 (Osborne
Reynolds Centenary Volume)

[13] Savill A M 1990 A synthesis of T3 test case predictions Numerical


Simulation of Unsteady Flows and Transition to Turbulence ed O
Pironneauet al (Cambridge: Cambridge University Press) Savill A M 1996
Turbulence and Transition Modellinged M Hallbacket al (Dordrecht: Kluwer
Academic) ch 6

[14] Schubauer G B and Klebanoff P S 1955 Contribution to the mechanism of


boundary-layer transition NACA T. N. 3489

[15] Suzen Y B and Huang P G 2000 Modeling of flow transition using an


intermittency transport equation J. Fluids Eng. 122 27384

[16] Tsuei H-S and Perot J B 2000 Turbomachinery predictions using the
turbulent potential model AIAA 2000 0135 (Reno, Nevada, January 2000)

[17] Warsi Z U A 1999 Fluid Dynamics Theoretical and Computational


Approaches (Boca Raton, FL: Chemical Rubber Company)

[18] Well C S 1967 Effects of free stream turbulence on boundary layer transition
AIAA J. 5 1724

[19] White F M 1991 Viscous Fluid Flow (New York: McGraw-Hill)

79
[20] Van Driest E R and Blumer C B 1963 Boundary layer transition, free stream
turbulence, and pressure gradient effects AIAA J. 1 13036

[21] Zhang X and Perot J B 2000 Unsteady flow prediction of turbulent flow
around a triangular cylinder FEDSM 200011172 (Boston, MA, June 2000)

[22] E. Aurell, G. Boffetta, A. Crisanti, G. Paladin, and a. Vulpiani. Growth of


nonifinitesimal perturbations in turbulence. Physical review letters,
77(7):1262, 1996.

[23] S. Chen and X. Shan. High-resolution turbulent simulations using the


connection machine-2. Computers in Physics, 6(6) : 643-646, 1992.

[24] G. Comte-Bellot and S.Corrsin. The use of a contraction to improve the


isotropy of grid-generated turbulence. Journal of fluid mechanics, 25(4): 657-
682, 1966.

[25] G. Comte-Bellot and S. Corrsin. Simple eulerian time correlation of full-and


narrow-band velocity signals in grid generated, isotropic turbulence.
Journal of Fluid Mechanics, 48(2):273-337, 1971.

[26] A. Crisanti, M. Jensen, A. Vulpiani, and G. Paladin. Intermittency and


predictability in turbulence. Physical review leters, 70(2):166, 1993.

[27] P.A. Durbin and B.P. Reif. Statical theory and modeling for turbulent flows.
John Viley& Sons, 2011.

80

You might also like