16.08.2013 Views

The UMIST-N Near-Wall Treatment Applied to Periodic Channel Flow

The UMIST-N Near-Wall Treatment Applied to Periodic Channel Flow

The UMIST-N Near-Wall Treatment Applied to Periodic Channel Flow

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>The</strong> <strong>UMIST</strong>-N <strong>Near</strong>-<strong>Wall</strong><br />

<strong>Treatment</strong> <strong>Applied</strong> <strong>to</strong> <strong>Periodic</strong><br />

<strong>Channel</strong> <strong>Flow</strong><br />

Submitted for the Degree of<br />

Master of Philosophy<br />

by<br />

Bryn Richards<br />

Department of Mechanical, Aerospace<br />

and Manufacturing Engineering<br />

University of Manchester<br />

Institute of Science and Technology<br />

2005


Declaration<br />

No portion of the work referred <strong>to</strong> in this thesis has been submitted in sup-<br />

port of an application for another degree or qualification of this or any other<br />

university or other institution of learning.<br />

i


Acknowledgements<br />

I would like <strong>to</strong> thank my supervisor, Dr. A. P. Watkins for his guidance<br />

throughout this project and for his willingness <strong>to</strong> be interrupted and <strong>to</strong> invest<br />

time with me whenever I have sought his attention. His style as a supervisor<br />

allows flexibility, independence, and self-actualisation. Our discussions have<br />

generated possibilities more than they have limited them.<br />

I would like <strong>to</strong> thank Professor Dominique Laurence for his involvement. He<br />

has continually made connections that have identified further sources of input<br />

in<strong>to</strong> the project and further applicability <strong>to</strong> the work of others.<br />

Dr. T. J. Craft has offered access <strong>to</strong> his unparalleled power of explanation<br />

and keen investigative ability whenever I have approached his open door.<br />

Dr. Simon Gant has offered patient and careful explanations on innumerable<br />

occasions. I greatly admire the enthusiasm, stewardship, and excellence that<br />

he brings <strong>to</strong> every facet of his professional life.<br />

Many others throughout the department have acted as sources of inspiration<br />

and support. I will continue <strong>to</strong> benefit from the diverse perspectives and<br />

experiences of the many interesting people whom I have had the good fortune<br />

<strong>to</strong> meet during my time at <strong>UMIST</strong>.<br />

I gratefully acknowledge the support of the DESIDER project in partially<br />

funding this research.<br />

ii


Abstract<br />

This thesis assesses the performance of the <strong>UMIST</strong>-N subgrid near-wall treat-<br />

ment when applied <strong>to</strong> periodic flow. <strong>The</strong> thesis also assesses the use of the<br />

k-ω turbulence model with the subgrid approach. Based on this work, the<br />

approach appears <strong>to</strong> be applicable <strong>to</strong> time-variant flow. Further research is<br />

required <strong>to</strong> improve the implementation of the k-ω model.<br />

<strong>UMIST</strong>-N solves simplified transport equations near a solid boundary at a<br />

lower computational cost than that of a low-Reynolds-number treatment.<br />

In applying the method <strong>to</strong> periodic channel flow, none of the approaches<br />

considered performed in an exemplary manner, but the subgrid exhibited no<br />

apparent failing when compared <strong>to</strong> the low-Reynolds-number results. It did<br />

offer an improvement upon the logarithmic law of the wall.<br />

This work highlights a numerical difficulty in applying the subgrid solution<br />

as a boundary layer <strong>to</strong> the main grid when the k-ω model is used. This has<br />

manifest itself in the results as an enhanced propensity <strong>to</strong>ward a discontinuity<br />

in calculated profiles at the subgrid / main grid interface under certain flow<br />

conditions.<br />

A range of steady channel flow data has been compiled and presented in this<br />

thesis. Empirical correlations are offered which identify general tendencies<br />

in the data and may provide a useful <strong>to</strong>ol for researchers engaged in the<br />

computation of flows.<br />

iii


Contents<br />

1 Introduction & Literature Survey 1<br />

1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.2 Turbulence Models . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.3 <strong>Wall</strong> Functions & the Subgrid Approach . . . . . . . . . . . . 7<br />

1.4 Relevance <strong>to</strong> Large Eddy Simulation . . . . . . . . . . . . . . 11<br />

1.5 <strong>Periodic</strong> <strong>Flow</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.6 Study Objectives . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.7 <strong>The</strong>sis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2 Turbulence Models 17<br />

2.1 Reynolds Averaging . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2 <strong>The</strong> k-ε Model . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.1 <strong>The</strong> Low-Reynolds-Number k-ε Model . . . . . . . . . 22<br />

2.2.2 Yap Correction . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.3 <strong>The</strong> k-ω Model . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.4 <strong>The</strong> Logarithmic Law of the <strong>Wall</strong> . . . . . . . . . . . . . . . . 26<br />

3 <strong>Channel</strong> <strong>Flow</strong> 31<br />

3.1 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.1.1 <strong>The</strong> k-ε Model . . . . . . . . . . . . . . . . . . . . . . 33<br />

iv


3.1.2 <strong>The</strong> k-ω Model . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.2 <strong>Flow</strong> Characterisation . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.3 Nondimensionalisation . . . . . . . . . . . . . . . . . . . . . . 36<br />

3.4 Steady <strong>Channel</strong> <strong>Flow</strong> Data . . . . . . . . . . . . . . . . . . . . 37<br />

3.4.1 Empirical Profile for U + . . . . . . . . . . . . . . . . . 39<br />

3.4.2 Empirical Profile for − 〈uv〉 + . . . . . . . . . . . . . . 39<br />

3.4.3 Empirical Profile for k + . . . . . . . . . . . . . . . . . 40<br />

3.4.4 Empirical Profile for 〈u 2 〉 + . . . . . . . . . . . . . . . . 41<br />

3.4.5 Empirical Profile for 〈v 2 〉 + . . . . . . . . . . . . . . . . 42<br />

3.4.6 <strong>Near</strong>-<strong>Wall</strong> Behaviour . . . . . . . . . . . . . . . . . . . 43<br />

3.5 Local Nondimensionalisation . . . . . . . . . . . . . . . . . . . 45<br />

4 Numerical Implementation 48<br />

4.1 <strong>The</strong> Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

4.2 Volume Integral Form . . . . . . . . . . . . . . . . . . . . . . 50<br />

4.3 Discretisation . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

4.4 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 54<br />

4.4.1 <strong>Wall</strong> Boundaries on k-ε . . . . . . . . . . . . . . . . . . 55<br />

4.4.2 <strong>Wall</strong> Boundaries on k-ω . . . . . . . . . . . . . . . . . 55<br />

4.4.3 <strong>The</strong> Logarithmic Law of the <strong>Wall</strong> . . . . . . . . . . . . 56<br />

4.4.4 <strong>The</strong> Subgrid Approach . . . . . . . . . . . . . . . . . . 60<br />

4.5 Under-Relaxation . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

5 Results 66<br />

5.1 Steady <strong>Flow</strong> Results . . . . . . . . . . . . . . . . . . . . . . . 68<br />

5.2 Prescribed <strong>Periodic</strong> Pressure Gradient . . . . . . . . . . . . . 71<br />

5.3 Prescribed <strong>Periodic</strong> Bulk <strong>Flow</strong> Rate . . . . . . . . . . . . . . . 75<br />

6 Conclusions & Suggestions for Future Work 83<br />

v


List of Figures<br />

2.1 <strong>The</strong> log-law compared <strong>to</strong> experiments in channel flow . . . . . 30<br />

4.1 <strong>The</strong> low-Reynolds-number grid . . . . . . . . . . . . . . . . . 49<br />

4.2 <strong>The</strong> high-Reynolds-number grid . . . . . . . . . . . . . . . . . 57<br />

4.3 <strong>The</strong> subgrid mesh, adapted from Gant [21] . . . . . . . . . . . 60<br />

3.1 Reichardt’s law <strong>to</strong> estimate U + . . . . . . . . . . . . . . . . . 100<br />

3.2 Reichardt’s law applied <strong>to</strong> − 〈uv〉 + . . . . . . . . . . . . . . . 101<br />

3.3 A revised profile for − 〈uv〉 + . . . . . . . . . . . . . . . . . . . 102<br />

3.4 A profile for k + . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

3.5 A profile for 〈uu〉 + . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

3.6 A profile for 〈vv〉 + . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

3.7 <strong>Near</strong>-wall behaviour of flow parameters . . . . . . . . . . . . . 106<br />

3.8 y ∗ vs. y + . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

3.9 U ∗ superimposed on U + . . . . . . . . . . . . . . . . . . . . . 108<br />

3.10 y ∗ v2 vs. y + . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109<br />

3.11 U ∗ v2 superimposed on U + . . . . . . . . . . . . . . . . . . . . . 110<br />

5.1 〈U〉 + vs y/δ in the steady flow case . . . . . . . . . . . . . . . 111<br />

5.2 k + vs y/δ in the steady flow case . . . . . . . . . . . . . . . . 112<br />

5.3 〈U〉 + vs y + in the steady flow case . . . . . . . . . . . . . . . 113<br />

5.4 k + vs y + in the steady flow case . . . . . . . . . . . . . . . . . 114<br />

vi


5.5 Bulk flow variation in the periodic pressure case . . . . . . . . 115<br />

5.6 <strong>Wall</strong> shear stress variation in the periodic pressure case . . . . 116<br />

5.7 Variables with time at y/δ = 0.1 (prescribed<br />

5.8 Variables with time at y/δ = 0.2 (prescribed<br />

5.9 Variables with time at y/δ = 0.5 (prescribed<br />

5.10 Variables with time at y/δ = 0.9 (prescribed<br />

5.11 〈U〉 vs y/δ snapshots through time (prescribed<br />

5.12 〈U〉 vs y/δ snapshots through time (prescribed<br />

5.13 k vs y/δ snapshots through time (prescribed<br />

5.14 k vs y/δ snapshots through time (prescribed<br />

5.15 〈U〉 vs y + snapshots through time (prescribed<br />

5.16 〈U〉 vs y + snapshots through time (prescribed<br />

5.17 k vs y + snapshots through time (prescribed<br />

5.18 k vs y + snapshots through time (prescribed<br />

∂〈P 〉<br />

) . . . . . . . 117<br />

∂x<br />

∂〈P 〉<br />

) . . . . . . . 118<br />

∂x<br />

∂〈P 〉<br />

) . . . . . . . 119<br />

∂x<br />

∂〈P 〉<br />

) . . . . . . . 120<br />

∂x<br />

∂〈P 〉<br />

) - part 1 . 121<br />

∂x<br />

∂〈P 〉<br />

) - part 2 . 122<br />

∂x<br />

∂〈P 〉<br />

) - part 1 . . 123<br />

∂x<br />

∂〈P 〉<br />

) - part 2 . . 124<br />

∂x<br />

∂〈P 〉<br />

) - part 1 . 125<br />

∂x<br />

∂〈P 〉<br />

) - part 2 . 126<br />

∂x<br />

∂〈P 〉<br />

) - part 1 . . . 127<br />

∂x<br />

∂〈P 〉<br />

) - part 2 . . . 128<br />

∂x<br />

5.19 Pressure variation in the periodic bulk flow . . . . . . . . . . . 129<br />

5.20 <strong>Wall</strong> shear stress variation in the periodic bulk flow case . . . 130<br />

5.21 Variables with time at y/δ = 0.1 (prescribed U) . . . . . . . . 131<br />

5.22 Variables with time at y/δ = 0.2 (prescribed U) . . . . . . . . 132<br />

5.23 Variables with time at y/δ = 0.5 (prescribed U) . . . . . . . . 133<br />

5.24 Variables with time at y/δ = 0.9 (prescribed U) . . . . . . . . 134<br />

5.25 〈U〉 vs y/δ snapshots through time (prescribed U) - part 1 . . 135<br />

5.26 〈U〉 vs y/δ snapshots through time (prescribed U) - part 2 . . 136<br />

5.27 k vs y/δ snapshots through time (prescribed U) - part 1 . . . 137<br />

5.28 k vs y/δ snapshots through time (prescribed U) - part 2 . . . 138<br />

5.29 〈U〉 vs y + snapshots through time (prescribed U) - part 1 . . . 139<br />

5.30 〈U〉 vs y + snapshots through time (prescribed U) - part 2 . . . 140<br />

5.31 k vs y + snapshots through time (prescribed U) - part 1 . . . . 141<br />

vii


5.32 k vs y + snapshots through time (prescribed U) - part 2 . . . . 142<br />

viii


List of Tables<br />

2.1 Constants in the standard k-ε model [33] . . . . . . . . . . . . 22<br />

2.2 Constants in the Wilcox 1988 k-ω model [89] . . . . . . . . . . 26<br />

2.3 Log-law constants [83] . . . . . . . . . . . . . . . . . . . . . . 27<br />

2.4 <strong>Near</strong>-wall flow regimes adapted from [53] . . . . . . . . . . . . 28<br />

3.1 Qualitative behaviour of terms in Equation 3.41 . . . . . . . . 41<br />

3.2 Qualitative behaviour of terms in Equation 3.42 . . . . . . . . 42<br />

4.1 <strong>The</strong> notation for discretised values . . . . . . . . . . . . . . . 50<br />

4.2 Under-relaxation fac<strong>to</strong>rs used in the main grid . . . . . . . . . 65<br />

4.3 Under-relaxation fac<strong>to</strong>rs used in the subgrid . . . . . . . . . . 65<br />

5.1 Configurations of turbulence models . . . . . . . . . . . . . . . 67<br />

ix


Chapter 1<br />

Introduction & Literature<br />

Survey<br />

1.1 Background<br />

Computational Fluid Dynamics (CFD) is a field of study that seeks <strong>to</strong> simu-<br />

late and predict fluid flow using computers. CFD may occasionally be used<br />

<strong>to</strong> investigate the physical behaviour of flow, but the usual goal is <strong>to</strong> provide<br />

an input <strong>to</strong> engineering analysis and design. Following from this goal, the<br />

thrust of most CFD development is <strong>to</strong> continually reduce the extent of the<br />

tradeoff between the predictive accuracy and the computational affordability<br />

of CFD. <strong>The</strong> field of CFD has developed quickly in recent decades because<br />

of the rapid advance and increasing accessibility of computer technology.<br />

Having first made CFD possible, then practical as an engineering <strong>to</strong>ol, this<br />

ongoing advancement increasingly drives a trend <strong>to</strong>ward more complex cal-<br />

culations. Recognising this, researchers in CFD tend <strong>to</strong> direct their efforts<br />

<strong>to</strong>ward enhancing the predictive accuracy per unit cost of CFD treatments<br />

1


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 2<br />

that are ever more complex.<br />

<strong>The</strong> type of flow considered in this thesis is incompressible and New<strong>to</strong>nian.<br />

On scales that are several orders of magnitude larger than the molecular<br />

scale, the flow is governed by the Navier-S<strong>to</strong>kes equations. <strong>The</strong>se equations<br />

model a flow field as a continuum. CFD codes discretise the flow domain<br />

<strong>to</strong> allow variables <strong>to</strong> be represented numerically. <strong>The</strong> most common type of<br />

discretisation is the Finite Volume method, which is applied in this work. <strong>The</strong><br />

governing equations of fluid flow identify a large degree of interconnectedness<br />

between the various properties of a flow field. Because of this, most CFD<br />

codes use an iterative approach. This involves traversing the flow field and<br />

everywhere calculating flow parameters based on local information. <strong>The</strong><br />

solution converges as this process is repeated with updated local information.<br />

In general, a flow field is a complex system with mechanisms for internal<br />

feedback. Because of this, a flow field can exhibit chaotic behaviour. This<br />

means that, although fluid flow is deterministic, it may appear <strong>to</strong> include<br />

randomness because of its extreme physical complexity. A flow is classed as<br />

turbulent when it behaves chaotically. This is a subjective definition, and<br />

there does exist ‘transitional flow’ where the appropriateness of classing the<br />

flow as turbulent is unclear. However, most flows of relevance <strong>to</strong> engineering<br />

applications are clearly turbulent, and most CFD work is aimed at predicting<br />

the behaviour of turbulent flow. Most commonly, this is accomplished by<br />

the adoption of a turbulence model that makes use of certain assumptions <strong>to</strong><br />

approximate flow behaviour using more tractable equations. This precludes<br />

the possibility of predicting every feature of the flow, but the parameters<br />

providing the greatest engineering relevance may be estimated at a greatly<br />

reduced computational cost.


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 3<br />

One approach <strong>to</strong> turbulence modelling is the Reynolds Averaged Navier-<br />

S<strong>to</strong>kes (RANS) type. RANS models use time- or ensemble-averaged Navier-<br />

S<strong>to</strong>kes equations <strong>to</strong> calculate the mean values of flow parameters. 1 <strong>The</strong><br />

fluctuating components of these parameters are modelled, rather than being<br />

fully resolved. Conceptually, this amounts <strong>to</strong> solving a flow as though it<br />

were laminar, but with the addition of modelled turbulence superimposed<br />

over the bulk flow behaviour. This modelled turbulence affects the bulk flow<br />

according <strong>to</strong> the details of the model.<br />

<strong>The</strong> most popular RANS models are the Eddy-Viscosity Models (EVMs).<br />

EVMs model the affect of turbulence on bulk flow via the concept of turbulent<br />

viscosity. Local turbulence is presumed <strong>to</strong> manifest itself as an increase in<br />

the effective viscosity of the fluid. <strong>The</strong> physical justification for this concept<br />

is that turbulence entails greater interaction between fluid particles. This<br />

leads <strong>to</strong> a greater exchange of energy between adjacent parcels of fluid. In<br />

terms of the momentum equations, this interaction produces an effect that<br />

is analogous <strong>to</strong> increased viscosity.<br />

<strong>The</strong> EVM may be extended by the incorporation of non-linear terms [77, 14]<br />

This is the non-linear type of EVM. An alternative <strong>to</strong> EVMs is presented<br />

by the Reynolds Stress Transport models [69, 15]. Here, the products and<br />

squares of root-mean-square velocity fluctuations (Reynolds stresses) are cal-<br />

culated as being convected and diffused through the flow field according <strong>to</strong><br />

their own transport equations. This thesis deals with the application of linear<br />

EVM RANS.<br />

1 Ensemble averaging is associated with periodic flow problems, and time averaging is<br />

associated with steady problems.


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 4<br />

Another type of CFD is Direct Numerical Simulation (DNS) [30]. DNS in-<br />

volves numerical approximations <strong>to</strong> discretize the flow field, but no turbu-<br />

lence model is used. Instead, DNS resolves all the quantities associated<br />

with a flow, including small-scale turbulent fluctuations. DNS is very com-<br />

putationally expensive, and is only employed for problems involving simple<br />

geometries and low Reynolds numbers. Although there is a potential for<br />

error in DNS studies, DNS results are usually taken <strong>to</strong> represent true fluid<br />

behaviour, when compared against modelled results. Because DNS provides<br />

more complete information than do experiments, DNS data are often used<br />

<strong>to</strong> validate turbulence models.<br />

A popular test case for CFD is channel flow. <strong>The</strong> geometrical simplicity<br />

of channel flow allows solutions <strong>to</strong> be achieved at minimal computational<br />

expense and also reduces the complexity of the CFD code, minimizing the<br />

potential for errors in coding. This thesis deals with channel flow. In addi-<br />

tion <strong>to</strong> the steady-flow case, a variable pressure gradient is applied <strong>to</strong> test<br />

the impact of periodic fluctuations on the solution method. <strong>The</strong> solution of<br />

channel flow leads <strong>to</strong> a system of equations that is parabolic, meaning that<br />

they may be solved using only local and upstream information. In the case<br />

of steady flow, the flow field is also statistically stationary, meaning that flow<br />

statistics such as mean flow and turbulence levels are invariant with time.<br />

In the case where a variable pressure gradient is applied, the flow field is<br />

not strictly stationary, but a converged solution produces results which are<br />

statistically stationary with respect <strong>to</strong> a given phase angle of pressure gradi-<br />

ent fluctuation. Thus, ensemble averaging of the Navier-S<strong>to</strong>kes equations is<br />

used.


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 5<br />

1.2 Turbulence Models<br />

RANS models are based upon the idea of filtering turbulence from the gov-<br />

erning equations of a flow so that it may be treated separately. This is due<br />

<strong>to</strong> Reynolds [59]. In EVMs, the problem of knowing the effect of turbulence<br />

on the mean flow is made tractable by employing the idea of a turbulent<br />

viscosity, based on the work of Boussinesq [6]. Turbulent viscosity may be<br />

calculated in a number of different ways.<br />

<strong>The</strong> simplest approach is <strong>to</strong> specify the turbulent viscosity at a given location<br />

based on known local quantities. Models based on this approach are called<br />

‘algebraic models’. Taylor [80] and Prandtl [54] have proposed an algebraic<br />

model in which turbulent viscosity is calculated as a function of a length scale<br />

and a local mean velocity gradient. <strong>The</strong> length scale is specified as a function<br />

of wall-distance. Unfortunately, an appropriate length scale can be difficult<br />

<strong>to</strong> obtain in complex geometries. Another limitation of algebraic models is<br />

that the dependence of turbulent viscosity on a local mean velocity gradient<br />

is unrealistic in many types of flow.<br />

To remove the latter limitation of algebraic models, ‘one-equation’ models<br />

track an additional turbulent quantity through the solution of an additional<br />

transport equation. <strong>The</strong> most popular choice of the additional quantity is<br />

the turbulent kinetic energy per unit mass, k (usually referred <strong>to</strong> simply as<br />

turbulent kinetic energy). Prandtl [55] proposed this approach. Spalart et<br />

al. [73] have proposed directly solving a transport equation for turbulent<br />

viscosity. A limitation that is common <strong>to</strong> all one-equation models is that a<br />

length scale is still required <strong>to</strong> fully specify the modelled flow.


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 6<br />

To remove the dependence on a length scale, ‘two-equation’ models intro-<br />

duce an additional transported quantity. Various choices exist for the second<br />

quantity. Kolmogorov [31] suggested k 1<br />

2 /l, where l is the mixing length. This<br />

quantity has later been dubbed ω. Chou [9] proposed modelling the rate of<br />

dissipation of turbulence, ε. In terms of dimensional analysis, this amounts<br />

<strong>to</strong> modelling k 3<br />

2 /l. However, it may be argued that turbulent dissipation,<br />

which involves the conversion of turbulent energy <strong>to</strong> heat, takes place on a<br />

physically much smaller scale than turbulence itself, so that ε and k 3<br />

2 /l may<br />

not be the same thing [91].<br />

Other choices for the transported quantities bear mentioning. Rotta [63]<br />

and Spalding [74] proposed l as the second quantity. Rotta [64], Rodi &<br />

Spalding [62], and Ng & Spalding [46] proposed models using the product kl.<br />

Spalding [75], Wilcox & Rubesin [88], and Robinson et al. [61] have proposed<br />

ω 2 . Coakley [10] proposed k 1<br />

2 and ω as the two transported quantities. <strong>The</strong>se<br />

two-equation models are compared and discussed by Wilcox [91].<br />

<strong>The</strong> most popular choice of two-equation model is the k-ε model of Launder<br />

& Sharma [33]. This is a revised tuning of the k-ε model of Jones & Launder<br />

[27]. <strong>The</strong>se papers follow the work of Chou [9], Davidov [19], and Harlow &<br />

Nakayama [23]. <strong>The</strong> work of Jones & Launder introduced a modification <strong>to</strong> k-<br />

ε modelling that allowed the model <strong>to</strong> be applied in the near-wall region. <strong>The</strong><br />

model of Launder & Sharma is commonly called the ‘standard’ k-ε model.<br />

Yap [92] improved the performance of the standard k-ε model in impinging<br />

and recirculating flow. Yap’s modification, the ‘Yap correction’ is a popular<br />

addition <strong>to</strong> the model.<br />

<strong>The</strong> k-ω model is a popular alternative <strong>to</strong> the k-ε model. Following the


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 7<br />

work of Kolmogorov [31] and Saffman [65], Wilcox [89] proposed the most<br />

well-known k-ω model. <strong>The</strong> model does not require the same damping terms<br />

employed by the standard k-ε model in order <strong>to</strong> be effective in the near-wall<br />

region, but boundary conditions are more difficult <strong>to</strong> apply. <strong>The</strong> principal<br />

difficulty associated with the k-ω model is in specifying a value for ω in free-<br />

stream turbulence [40]. Speziale et al. [76], Menter [41], Peng et al. [52],<br />

and Wilcox [90] have proposed modifications <strong>to</strong> the k-ω that further improve<br />

its performance at the cost of adding much the same degree of complexity<br />

found in the standard k-ε model.<br />

<strong>The</strong> SST 2 model of Menter [42] may be thought of as a hybrid k-ε / k-ω<br />

approach. It employes a k-ω model near solid boundaries and a k-ε model<br />

elsewhere. <strong>The</strong> SST model is appealing because the principal strength of<br />

the k-ω model is its simplicity and relative accuracy in the near-wall region,<br />

while the k-ε model is generally more effective in free-stream flow. <strong>The</strong> SST is<br />

implemented by the use of a blending function <strong>to</strong> provide a smooth transition<br />

between the two models.<br />

1.3 <strong>Wall</strong> Functions & the Subgrid Approach<br />

<strong>Near</strong> solid boundaries, the gradients of turbulence quantities become large.<br />

Numerically, this means that greater s<strong>to</strong>rage and computational demands are<br />

placed on a CFD code that performs calculations in the near-wall region. To<br />

avoid this, wall functions are often employed. A wall function is a solution<br />

method that provides a means of characterising turbulence at some point,<br />

2 Shear Stress Transport


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 8<br />

P away from the wall. <strong>The</strong> calculations performed by the wall function<br />

are simpler than applying a full CFD approach between the wall and the<br />

point P . Thus, a computational savings is achieved. <strong>The</strong> CFD code then<br />

calculates the remaining flow field, taking the wall-function output at point P<br />

as a boundary condition, replacing the actual wall boundary condition. <strong>The</strong><br />

disadvantage of wall functions is that they cannot provide the same level of<br />

accuracy as a full CFD treatment near the wall, except in certain flows for<br />

which they were explicitly designed. 3<br />

<strong>The</strong> most common type of wall function is that which relies on the ‘logarith-<br />

mic law of the wall’. This empirical equation relates velocity <strong>to</strong> wall-normal<br />

distance within a certain near-wall region. Launder & Spalding [34] have<br />

produced a log-law-based wall-function that is a popular default in many<br />

industrial applications. In addition <strong>to</strong> offering a means of calculating near-<br />

wall velocity, turbulent kinetic energy, k is calculated by tracking the rates<br />

of production and dissipation of this quantity, averaged analytically over the<br />

near-wall region. This wall function acts as a one-equation model in the<br />

sense that it only resolves production and dissipation of one quantity. When<br />

used as a boundary condition on two-equation models, the second turbulence<br />

quantity is calculated from k and the length scale. Launder & Spalding offer<br />

a means of calculating ε. Wilcox [91] offers a similar wall function for the<br />

k-ω model.<br />

Various other wall treatments exist. <strong>The</strong>se may be thought of as falling on<br />

a spectrum between a costly but relatively accurate full CFD solution and,<br />

as the other polar extreme, a set of simple algebraic expressions resulting<br />

3 <strong>The</strong> great majority of wall functions have been designed for steady channel flow,<br />

including the wall function of Launder & Spalding [34] employed in this work.


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 9<br />

from an analytical treatment of some presumed near-wall behaviour. Be-<br />

tween these two extremes, various alternatives exist. Broadly, these may be<br />

classified as either attempting <strong>to</strong> extend the applicability of standard wall<br />

functions through the incorporation of more complex mathematics and more<br />

accurate empirical and theoretical information or attempting <strong>to</strong> simplify CFD<br />

near the wall <strong>to</strong> achieve accurate near-wall solutions at less computational<br />

expense than the full CFD treatment.<br />

In the realm of extending the applicability of analytical wall functions, Amano<br />

[3] presented a wall function that finds average production and destruction<br />

of ε, in addition <strong>to</strong> k. Thus, the use of the length scale <strong>to</strong> calculate ε is<br />

avoided. Smith [72] and Craft et al. [17] have proposed wall functions using<br />

numerical, rather than analytical methods <strong>to</strong> obtain near-wall averaged val-<br />

ues within the wall function. This allows the use of more complex equations<br />

for turbulence quantities in the wall function, more closely approximating<br />

the RANS equations that would be solved by a full CFD treatment. Viegas<br />

& Rubesin [84] offered a wall function for compressible flow.<br />

Much progress has been made at <strong>UMIST</strong> in reducing the computational<br />

cost of CFD near solid boundaries. This approach may still be classified<br />

as a wall function treatment, in the sense that the main CFD code sees<br />

a ‘black box’ calculation scheme offering values for flow quantities in the<br />

near-wall region obtained at a reduced computational cost. However, the<br />

approach is not analytical, and instead may be described as using a simplified<br />

CFD calculation applied <strong>to</strong> the near-wall region. Thus, a potential exists<br />

for this simplified CFD approach <strong>to</strong> offer enhanced accuracy over other wall<br />

functions, although the implementation is generally more costly and complex.


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 10<br />

A precursor <strong>to</strong> simplified CFD wall functions is the PSL 4 approach of Ia-<br />

covides [25]. This is a modification <strong>to</strong> the full CFD near-wall solution that<br />

constrains the CFD code <strong>to</strong> ignore the coupling between velocity and pressure<br />

in the near-wall region. Instead, the static pressure distribution is taken as<br />

fixed in this region. This results in significant computational savings. Craft<br />

et al. [18] extended this approach by encapsulating the near-wall CFD cal-<br />

culation as an independent code using its own subgrid <strong>to</strong> perform simplified<br />

CFD calculations in the near-wall region. Thus, the simplified CFD approach<br />

was encapsulated as a wall function. This approach is called <strong>UMIST</strong>-N 5 . As<br />

with PSL, the assumption of constant pressure distribution results in equa-<br />

tions which are parabolic and may be solved using simplified methods. A<br />

momentum equation is not employed for wall-normal velocity. Instead, lo-<br />

cal conservation of mass is employed in the near-wall region <strong>to</strong> determine<br />

wall-normal velocities from wall-parallel velocities.<br />

<strong>UMIST</strong>-N offers a simplified near-wall treatment, relative <strong>to</strong> a full CFD solu-<br />

tion, but it resolves a two-equation turbulence model near the wall. <strong>UMIST</strong>-<br />

N returns near-wall averaged production and destruction of two turbulence<br />

parameters <strong>to</strong> the main grid, eliminating the use of a length scale. <strong>The</strong><br />

development of <strong>UMIST</strong>-N is detailed in the PhD thesis of Gant [21]. <strong>The</strong><br />

k-ε model has been employed in <strong>UMIST</strong>-N, including the non-linear EVM<br />

k-ε model of Craft et al. [14]. <strong>UMIST</strong>-N has been applied <strong>to</strong> channel flow,<br />

impinging jet, spinning disk, and Ahmed body flow, and has been adapted<br />

<strong>to</strong> non-orthogonal coordinate systems [16, 21]. All flows considered thus far<br />

have been steady in time.<br />

4 Parabolic Sub-Layer<br />

5 Unified Modelling through Integrated Sublayer <strong>Treatment</strong> - a N umerical approach


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 11<br />

1.4 Relevance <strong>to</strong> Large Eddy Simulation<br />

Large Eddy Simulation (LES) is a modelling approach that allows large-<br />

scale turbulent fluctuations <strong>to</strong> remain represented within the Navier-S<strong>to</strong>kes<br />

equations while small-scale turbulent fluctuations are filtered out statistically<br />

and treated separately. Thus it can be thought of as offering a compromise<br />

between DNS and RANS. <strong>The</strong> LES approach involves the complexities of<br />

resolving large-scale turbulence, of modelling small-scale turbulence, and of<br />

handling the interaction between these two scales of turbulence. However,<br />

LES provides a potential for greater predictive accuracy than any RANS<br />

method.<br />

LES and RANS approaches are far from alien, and hybrid calculations have<br />

been undertaken. Labourasse & Sagaut [32] have run LES within an overall<br />

RANS calculation. This provided a solution that exhibited the robustness<br />

of a RANS method with some additional accuracy derived from the use of<br />

LES. Quéméré et al. [56] have run RANS and LES calculations alongside one<br />

another in different zones within a flow domain. <strong>The</strong>se hybrid investigations<br />

highlight the complimentary strengths of RANS and LES in some flows.<br />

LES and RANS face analogous tradeoffs in the treatment of flow near solid<br />

boundaries. Performing a detailed LES calculation <strong>to</strong> resolve the turbulence<br />

near a wall is very computationally expensive. In most LES calculations,<br />

a wall function based on the logarithmic law of the wall is used <strong>to</strong> specify<br />

boundary conditions at a finite distance away from the wall. Balaras et al. [5]<br />

improved upon this by obtaining wall shear stress from a near-wall subgrid<br />

result within an LES calculation. <strong>The</strong> subgrid employed an algebraic model<br />

<strong>to</strong> obtain wall-parallel velocity and thus obtain a wall shear stress <strong>to</strong> act as


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 12<br />

a boundary condition on the LES solution.<br />

<strong>The</strong> <strong>UMIST</strong>-N approach entails a more complex calculation than that used<br />

by Balaras et al. and may offer improved results if used in LES. Certainly,<br />

a potential exists for <strong>UMIST</strong>-N <strong>to</strong> provide value in LES calculations. In<br />

the future, advanced wall functions may help <strong>to</strong> alleviate the well-known<br />

challenge in LES of adequately resolving the grid in regions where turbulence<br />

is generated by ‘driving mechanisms’.<br />

Because LES calculations resolve large-scale turbulent structures, any local<br />

region of an LES calculation space can be exposed <strong>to</strong> rapid fluctuations as<br />

eddies shift and move through the flow field. <strong>The</strong>refore, the application of<br />

<strong>UMIST</strong>-N <strong>to</strong> periodic flow represents an important first step in assessing its<br />

suitability <strong>to</strong> LES.<br />

1.5 <strong>Periodic</strong> <strong>Flow</strong><br />

<strong>Periodic</strong> flow refers <strong>to</strong> an arrangement in which the flow field varies smoothly<br />

and cyclicly as a function of time. This is sometimes referred <strong>to</strong> as oscillating<br />

flow. However, some researchers assert that the term ‘oscillating’ implies<br />

reciprocation, meaning that the direction of the bulk flow (and not merely<br />

its magnitude) is changing with time. This distinction is of questionable<br />

value. In this work, ‘periodic’ is used as the generic term. Furthermore,<br />

the term ‘periodic boundary conditions’ refers <strong>to</strong> boundary conditions that<br />

vary with time, driving periodic flow. This is distinct from the arrangement<br />

where values at an output boundary are copied <strong>to</strong> an input boundary in a<br />

loop. This type of boundary condition is also termed ‘periodic’, but these do


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 13<br />

not appear in the present work.<br />

<strong>Periodic</strong> flow provides an interesting test case, because it offers insight in<strong>to</strong><br />

some of the more illusive physical behaviours of a flow even in relatively<br />

simple geometries. Various experimentalists have investigated periodic flow<br />

in pipe and channel geometries. Sarpkaya [66] investigated periodically- and<br />

randomly-pulsed flow in a pipe from the standpoint of understanding the<br />

conditions under-which the flow transitioned <strong>to</strong> turbulence. Ohmi & Iguchi<br />

[48] performed an investigation in a similar vein, and Ohmi et al. [49, 50] also<br />

investigated higher Reynolds numbers. Tu & Ramaprian [82, 57] performed<br />

widely-recognised experimental work in<strong>to</strong> periodic pipe flow at a rather high<br />

Reynolds number. More recently, Tardu et al. [79] have investigated periodic<br />

channel flow. Hino et al. [24] performed an investigation of periodic flow at<br />

approximately the same time as Tu & Ramaprian, but their use of a rectan-<br />

gular duct geometry may have impacted the receptiveness of the modelling<br />

community <strong>to</strong> their results.<br />

Other interesting studies bear mentioning. Shemer et al. [68] offered a com-<br />

parison between laminar and turbulent flows at the same Reynolds number<br />

and under the same periodic conditions. Sleath [71] and Jensen et al. [26]<br />

investigated the impact of various surface roughnesses. Siginer [70] investi-<br />

gated periodic pipe flow using a non-New<strong>to</strong>nian fluid. Scotti & Piomelli [67]<br />

have applied LES <strong>to</strong> periodic channel flow at a range of oscillation frequen-<br />

cies. Lee et al. [36] and Walther et al. [85] have investigated heat transfer<br />

in periodic flow. Lodahl et al. [38] have investigated periodic pipe flow with<br />

a periodic applied electric current.<br />

Furthermore, internal combustion engine experiments offer insight in<strong>to</strong> pe-


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 14<br />

riodic flow. Ahmadi-Befrui et al. [2] published mean velocity readings at<br />

various locations within a cylinder through a pis<strong>to</strong>n stroke. Tabaczynski [78]<br />

has investigated reciprocating flow in engines. Also, the Society of Au<strong>to</strong>mo-<br />

tive Engineers has published an extensive collection of results.<br />

One very notable DNS study of periodic flow is that of Kawamura & Homma<br />

[29], which investigates channel flow at a low Reynolds number driven by a<br />

periodic pressure gradient. <strong>The</strong> periodic flow results presented in this thesis<br />

are compared against the DNS results of Kawamura & Homma.<br />

Various researcher have investigated the use of turbulence models in predict-<br />

ing periodic flow. Cot<strong>to</strong>n & Ismael [12] applied the standard k-ε model <strong>to</strong><br />

periodic pipe flow. Cot<strong>to</strong>n et al. [13] and Addad [1] have investigated the use<br />

of Reynolds Stress Transport models in predicting periodic flow. Watkins [86]<br />

produced an early investigation in<strong>to</strong> the use of CFD in internal combustion<br />

engines, and various other studies have followed [60, 35]. Nai<strong>to</strong>h & Kuwahara<br />

[45] applied LES <strong>to</strong> engine flow.<br />

<strong>The</strong> relevance of wall-functions <strong>to</strong> periodic flow has received less attention.<br />

<strong>The</strong> <strong>UMIST</strong>-N subgrid wall function has not been applied <strong>to</strong> periodic flow<br />

prior <strong>to</strong> this work. Standard wall functions are generally applied <strong>to</strong> internal<br />

combustion engine studies.<br />

1.6 Study Objectives<br />

<strong>The</strong> primary objective of this study is <strong>to</strong> assess the applicability of the<br />

<strong>UMIST</strong>-N subgrid near-wall treatment <strong>to</strong> a periodic flow problem, relative <strong>to</strong>


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 15<br />

other popular approaches. <strong>The</strong> secondary objective is <strong>to</strong> experiment with the<br />

use of the k-ω model in the subgrid solution scheme. This work represents<br />

the first application of the <strong>UMIST</strong>-N approach <strong>to</strong> time-variant flow and the<br />

first use of a k-ω model within the subgrid calculation.<br />

<strong>The</strong>se objectives are met through the analysis of the logarithmic law of the<br />

wall, the standard low-Reynolds number k-ε treatment, a k-ε subgrid treat-<br />

ment, and a k-ω subgrid treatment in steady channel flow and in periodically<br />

variable channel flow. <strong>The</strong> results of this study are compared against the de-<br />

tailed DNS data of Kim et al. [30] in the case of steady flow and Kawamura<br />

& Homma [29] in the case of periodic channel flow.<br />

A tertiary objective that arose over the course of the project was <strong>to</strong> inves-<br />

tigate the study of channel flow in general, so as <strong>to</strong> provide some input <strong>to</strong><br />

other CFD efforts that make use of channel flow data. This objective is met<br />

through a detailed background discussion of channel flow, a compilation of<br />

some useful experimental and DNS results concerning channel flow, and the<br />

revision and presentation of a set of analytical profiles <strong>to</strong> characterise the ex-<br />

pected behaviours of flow parameters in a channel as a function of Reynolds<br />

number.<br />

1.7 <strong>The</strong>sis Outline<br />

<strong>The</strong> various turbulence models employed in this work are discussed in Chap-<br />

ter 2. This includes a more detailed discussion of the RANS approach and<br />

a presentation of the k-ε and k-ω models. <strong>The</strong> differences between the high-<br />

and low-Reynolds-number k-ε models are highlighted. <strong>The</strong> logarithmic law


CHAPTER 1. INTRODUCTION & LITERATURE SURVEY 16<br />

of the wall is introduced.<br />

Chapter 3 contains a detailed discussion of channel flow. <strong>The</strong> governing equa-<br />

tions associated with the various turbulence models are adapted <strong>to</strong> channel<br />

flow. <strong>The</strong> chapter also presents various experimental and DNS channel flow<br />

results arising from other studies. Correlations are presented <strong>to</strong> identify<br />

trends in the data that may be useful for the design and testing of CFD<br />

codes. Alternative nondimensionalisation schemes are also investigated.<br />

Chapter 4 contains a discussion of the numerical implementation of the tur-<br />

bulence models. This includes a discussion of the Finite Volume method and<br />

the particular grid arrangements used in this work. Boundary conditions are<br />

also discussed, including the implementation of the logarithmic law of the<br />

wall. Since the novelty of <strong>UMIST</strong>-N is essentially in its numerical treatment<br />

of the boundary layer problem, the approach is detailed in this chapter.<br />

Chapter 5 presents an analysis of the results computed for the various config-<br />

urations considered. This includes steady channel flow, periodic flow results<br />

compared <strong>to</strong> data with the same driving pressure gradient, and periodic flow<br />

results compared <strong>to</strong> data with the same bulk flow rate.<br />

Chapter 6 offers conclusions and suggestions for future work.


Chapter 2<br />

Turbulence Models<br />

<strong>The</strong> Reynolds Averaged Navier S<strong>to</strong>kes equations are obtained for continuity,<br />

conservation of momentum, and generic transport. <strong>The</strong> assumptions of the<br />

EVM are employed where appropriate. <strong>The</strong>n, specific transport equations<br />

are introduced for the turbulence parameters tracked by the k-ε and k-ω<br />

models. <strong>The</strong> logarithmic law of the wall is introduced.<br />

2.1 Reynolds Averaging<br />

Consider a flow field containing an incompressible fluid with constant proper-<br />

ties: density (ρ), dynamic viscosity (µ), and kinematic viscosity (ν = µ<br />

). In ρ<br />

Cartesian coordinates, the flow field extends in three orthogonal directions,<br />

x, y, and z. <strong>The</strong> flow field may vary in time, t. A pressure field, P (x, y, z, t)<br />

and a velocity field, U (x, y, z, t) = (U, V, W ) T are associated with the flow.<br />

This flow field is governed by continuity and the conservation of momentum.<br />

17


CHAPTER 2. TURBULENCE MODELS 18<br />

Furthermore, it is assumed that a passive scalar, having no effect on the fluid<br />

properties, may be convected and diffused through the flow field according<br />

<strong>to</strong> a transport equation.<br />

Neglecting body forces such as gravity, the Navier-S<strong>to</strong>kes equations governing<br />

incompressible, constant-property New<strong>to</strong>nian fluid flow in Cartesian coordi-<br />

nates are<br />

Continuity : ▽ · U = 0 (2.1)<br />

Momentum : DU<br />

Dt<br />

T ransport :<br />

1 = − ▽ P + ▽ · (ν ▽ U) (2.2)<br />

ρ<br />

DΦ<br />

Dt = ▽ · (γ ▽ Φ) + Sφ (2.3)<br />

where Φ (x, y, z, t) represents a scalar quantity transported within the flow<br />

field. γ is the diffusivity of Φ in the fluid. In turbulence modelling, γ is<br />

usually replaced with ν , where σ is the effective Prandtl number of Φ in the<br />

σ<br />

fluid. Sφ (x, y, z, t) in Equation 2.3 is a net source of Φ.<br />

RANS modelling involves time- or ensemble-averaged Navier-S<strong>to</strong>kes equa-<br />

tions. <strong>The</strong> aim of Reynolds averaging is <strong>to</strong> allow separate tracking of mean<br />

flow and turbulent fluctuations. Because time-dependent flows are studied<br />

in this thesis, ensemble averaging is appropriate. Conceptually, an ensemble<br />

average is the mean of the instantaneous values of a parameter through a<br />

large number of repeated experiments. Ensemble averaging is similar <strong>to</strong> time<br />

averaging in that it separates turbulent fluctuations from the bulk flow by de-<br />

composing the velocity field in<strong>to</strong> a mean component (〈U〉) and a fluctuating<br />

velocity component (u), such that U = 〈U〉 + u. 1 Similarly, Φ = 〈Φ〉 + φ.<br />

1 Other popular notations include an over-bar for mean velocity U and a prime on<br />

fluctuating velocity (u ′ ).


CHAPTER 2. TURBULENCE MODELS 19<br />

<strong>The</strong> RANS equations are<br />

D<br />

Dt<br />

Continuity : ▽ · 〈U〉 = 0 (2.4)<br />

Momentum :<br />

T ransport :<br />

D〈U〉<br />

Dt<br />

D〈Φ〉<br />

Dt<br />

= −1 ▽ 〈P 〉 + ▽ · (ν ▽ 〈U〉)<br />

ρ<br />

∂ 〈uu〉 ∂ 〈vu〉 ∂ 〈wu〉<br />

− − −<br />

∂x ∂y ∂z<br />

<br />

ν<br />

<br />

= ▽ · ▽ 〈Φ〉 + Sφ<br />

σ<br />

∂ 〈uφ〉 ∂ 〈vφ〉 ∂ 〈wφ〉<br />

− − −<br />

∂x ∂y ∂z<br />

may be referred <strong>to</strong> as the mean material derivative and is defined by<br />

D<br />

Dt<br />

(2.5)<br />

(2.6)<br />

∂<br />

≡ + 〈U〉 · ▽ (2.7)<br />

∂t<br />

It can be shown [53] that the mean material derivative is related <strong>to</strong> the<br />

material derivative according <strong>to</strong><br />

<br />

DΘ<br />

=<br />

Dt<br />

D 〈Θ〉<br />

Dt<br />

for a quantity Θ = 〈Θ〉 + θ.<br />

∂ ∂ ∂<br />

+ 〈uθ〉 + 〈vθ〉 + 〈wθ〉 (2.8)<br />

∂x ∂y ∂z<br />

For convenience, Equation 2.5 may be re-written as<br />

D 〈U i〉 ∂ 〈P 〉<br />

= −1 +<br />

Dt ρ ∂xi ∂<br />

<br />

ν<br />

∂xj ∂ 〈U <br />

i〉<br />

−<br />

∂xj ∂ <br />

ujui ∂xj<br />

(2.9)<br />

where each side of the equation is a vec<strong>to</strong>r in i comprised of terms that are<br />

summed in j.<br />

By decomposing velocities in<strong>to</strong> mean and fluctuating components, RANS<br />

models seek <strong>to</strong> treat separately the physical details of turbulence. <strong>The</strong> fluctu-<br />

ating velocity components which are associated with turbulence are therefore


CHAPTER 2. TURBULENCE MODELS 20<br />

considered unknown, and must be modelled. <strong>The</strong> EVM models the unknown<br />

− <br />

ujui term in Equation 2.9 by<br />

− <br />

∂ 〈U<br />

ujui =<br />

i〉<br />

νt +<br />

∂xj ∂ <br />

U j<br />

−<br />

∂xi 2<br />

3 kδij (2.10)<br />

where k is turbulent kinetic energy. νt is turbulent viscosity, an additional<br />

viscosity arising as a result of turbulence. δij is the Kronecker delta, defined<br />

such that<br />

⎧<br />

⎨ 1 i = j<br />

δij =<br />

⎩ 0 i = j<br />

(2.11)<br />

Essentially, the EVM replaces ν with (ν + νt) in Equations 2.5 & 2.6. In<br />

summary, the governing equations of the RANS EVM are<br />

Continuity : ▽ · 〈U〉 = 0 (2.12)<br />

Momentum : D〈U〉<br />

Dt<br />

T ransport :<br />

= − 1<br />

ρ ▽ 〈P 〉 + ▽ · ((ν + νt) ▽ 〈U〉) (2.13)<br />

D〈Φ〉<br />

Dt = ▽ · ν+νt<br />

σ<br />

▽ 〈Φ〉 + Sφ<br />

(2.14)<br />

Thus the RANS and EVM method reduces the problem of calculating chaotic<br />

fluctuating velocity components <strong>to</strong> a problem of specifying an unknown local<br />

parameter, the turublent viscosity, νt. A variety of approaches exist for<br />

modelling νt. <strong>The</strong> approaches considered in this thesis are the k-ε model and<br />

the k-ω model.


CHAPTER 2. TURBULENCE MODELS 21<br />

2.2 <strong>The</strong> k-ε Model<br />

<strong>The</strong> high-Reynolds number version of the standard k-ε model of Launder &<br />

Sharma [33] defines turbulent viscosity, νt as<br />

2 k<br />

νt = Cµ<br />

ε<br />

(2.15)<br />

where k is turbulent kinetic energy, and ε is the rate of dissipation of k. Cµ<br />

is a constant given in Table 2.1.<br />

<strong>The</strong> transport equation for k is based on Equation 2.14 (with 〈Φ〉 = k). <strong>The</strong><br />

source term consists of production, Pk and dissipation, ε:<br />

<br />

Dk ν + νt<br />

= ▽ ·<br />

▽ k + Pk − ε (2.16)<br />

Dt σk<br />

Pk is<br />

Pk = −aij 〈Sij〉 (2.17)<br />

where the right hand side of the equation is summed over all permutations<br />

of i and j <strong>to</strong> obtain Pk.<br />

aij is defined as<br />

〈Sij〉 is defined as<br />

aij = u iu j<br />

〈Sij〉 ≡ 1<br />

<br />

∂ 〈U i〉<br />

2 ∂xj − 2<br />

3 kδij<br />

+ ∂ U j<br />

∂xi <br />

Applying the EVM (Equation 2.10) <strong>to</strong> Equation 2.17,<br />

(2.18)<br />

(2.19)<br />

Pk = 2νt 〈Sij〉 〈Sij〉 (2.20)


CHAPTER 2. TURBULENCE MODELS 22<br />

Equation 2.17 is an exact expression for Pk, within the assumptions of the<br />

RANS approach. Equation 2.20 is exact within the assumptions of the RANS<br />

EVM. However, ε is modelled. In k-ε models, ε is obtained through an addi-<br />

tional transport equation. In high-Reynolds-number version of the standard<br />

k-ε model of Launder & Sharma [33] the additional transport equation is<br />

<br />

Dε ν +<br />

<br />

νt<br />

ε<br />

2 ε<br />

= ▽ ·<br />

▽ ε + Cε1 Pk − Cε2<br />

(2.21)<br />

Dt σε<br />

k k<br />

<br />

production destruction<br />

<strong>The</strong> constants appearing in the standard k-ε model are given in Table 2.1.<br />

Table 2.1: Constants in the standard k-ε model [33]<br />

Cµ σk σε Cε1 Cε2<br />

0.09 1.0 1.3 1.44 1.92<br />

2.2.1 <strong>The</strong> Low-Reynolds-Number k-ε Model<br />

If the k-ε model is applied in the near-wall region, viscous corrections are<br />

required in order <strong>to</strong> produce reasonable results. Viscous corrections <strong>to</strong> the k-ε<br />

model involve two modifications. One is the incorporation of so-called viscous<br />

damping terms. Also, a transport equation is solved for a new parameter, ˜ε<br />

rather than ε.<br />

Because ε tends <strong>to</strong> a finite value at a wall, the wall boundary condition on<br />

ε is difficult <strong>to</strong> specify. <strong>The</strong>refore, the ε transport equation is replaced by a<br />

new transport equation of an alternative parameter, ˜ε. ˜ε is defined in such a


CHAPTER 2. TURBULENCE MODELS 23<br />

way that<br />

<strong>The</strong> difference between ˜ε and ε is defined as ˆε:<br />

From Equations 2.22 and 2.23, it follows that<br />

˜ε| y=0 = 0 (2.22)<br />

ε = ˜ε + ˆε (2.23)<br />

ˆε| y=0 = ε| y=0<br />

(2.24)<br />

<strong>The</strong> equation for νt is modified <strong>to</strong> use ˜ε for convenience. Also, damping<br />

terms exist in the equation for νt and the transport equation for ˜ε. <strong>The</strong>se<br />

damping terms are a function of the turbulent Reynolds number, ˜<br />

Ret. ˜<br />

Ret<br />

is a Reynolds number based on local turbulence quantities (in this case, k<br />

and ˜ε). <strong>The</strong> tilde is a reminder that, for convenience, the turbulent Reynolds<br />

number used in the low-Reynolds-number standard k-ε model uses ˜ε rather<br />

than ε. ˜ Ret is defined as<br />

˜Ret ≡ k2<br />

˜εν<br />

(2.25)<br />

In the low-Reynolds-number version of the standard k-ε model [33], the trans-<br />

port equation of ˜ε is<br />

<br />

D˜ε ν + νt<br />

= ▽ ·<br />

Dt σε<br />

where<br />

<br />

▽ ˜ε + Cε1f1<br />

E = 2ννt<br />

<br />

˜ε<br />

Pk − Cε2f2<br />

k<br />

∂ 2 〈U〉<br />

∂y 2<br />

2<br />

2 ˜ε<br />

+ E (2.26)<br />

k<br />

(2.27)<br />

f1 = 1 (2.28)<br />

−Ret ˜<br />

f2 = 1 − 0.3e 2<br />

(2.29)


CHAPTER 2. TURBULENCE MODELS 24<br />

ˆε is required in order <strong>to</strong> calculate ε for use in the k transport equation<br />

(Equation 2.16). ˆε is defined as<br />

ˆε = 2ν<br />

<br />

∂ √ 2 k<br />

∂y<br />

(2.30)<br />

Thus, ˆε does not require a transport equation, but may be solved from local<br />

quantities.<br />

νt is modified according <strong>to</strong><br />

where<br />

2.2.2 Yap Correction<br />

νt = Cµfµ<br />

fµ = exp<br />

⎡<br />

2 k<br />

˜ε<br />

⎣ −2.5<br />

<br />

1 + ˜ Ret<br />

50<br />

⎤<br />

(2.31)<br />

⎦<br />

(2.32)<br />

<strong>The</strong> ‘Yap correction’ refers <strong>to</strong> an additional source term in the ˜ε transport<br />

equation of the low-Reynolds-number k-ε model. <strong>The</strong> Yap correction was<br />

originally introduced by Yap [92] <strong>to</strong> improve the performance of the k-ε<br />

model in impinging and recirculating flows. Yap correction is often implicitly<br />

included in standard k-ε modelling, so it has been included in the present<br />

<strong>UMIST</strong>-N wall function for completeness.<br />

Using Yap correction, the ˜ε transport equation becomes<br />

2<br />

D˜ε ν + νt<br />

˜ε<br />

˜ε<br />

= ▽ ·<br />

▽ ˜ε + Cε1f1 Pk − Cε2f2 + E + Y (2.33)<br />

Dt σε<br />

k<br />

k<br />

with the additional Y source term defined as<br />

⎛⎡<br />

<br />

Y = max ⎝⎣0.83<br />

k 3<br />

2<br />

− 1<br />

2.5˜εy<br />

k 3<br />

2<br />

2.5˜εy<br />

2 2 ˜ε<br />

k<br />

⎤ ⎞<br />

⎦ , 0⎠<br />

(2.34)


CHAPTER 2. TURBULENCE MODELS 25<br />

y is the distance from the wall.<br />

2.3 <strong>The</strong> k-ω Model<br />

<strong>The</strong> k-ω model is like the k-ε model in being dubbed a two-equation model,<br />

because it demands the solution of two additional transport equations in<br />

order <strong>to</strong> characterise turbulence. Like most two-equation models, the k-ω<br />

model tracks turbulent kinetic energy, k. However, the dissipation rate, ε is<br />

not tracked using its own transport equation, as in the k-ε model. Instead, a<br />

transport equation is solved for ω. ω is sometimes referred <strong>to</strong> as the specific<br />

dissipation rate.<br />

<strong>The</strong> Wilcox 1988 k-ω model [89] uses the following transport equation for k:<br />

<br />

Dk ν + νt<br />

= ▽ ·<br />

▽ k + Pk − ωkβ<br />

Dt σkω<br />

∗<br />

(2.35)<br />

so that ω is defined as<br />

ω ≡ ε<br />

kβ ∗<br />

(2.36)<br />

Pk is defined in the same way as for the k-ε model, as shown in Equation<br />

2.20.<br />

<strong>The</strong> equation for νt becomes<br />

νt = γ ∗<br />

<strong>The</strong> transport equation for ω is<br />

<br />

Dω ν + νt<br />

= ▽ ·<br />

Dt σω<br />

<br />

k<br />

ω<br />

<br />

▽ ω + γ<br />

<br />

ω<br />

<br />

Pk − βω<br />

k<br />

2<br />

(2.37)<br />

(2.38)


CHAPTER 2. TURBULENCE MODELS 26<br />

<strong>The</strong> constants appearing in the Wilcox 1988 k-ω model are given in Table<br />

2.2.<br />

Table 2.2: Constants in the Wilcox 1988 k-ω model [89]<br />

β β ∗ γ γ ∗ σkω σω<br />

0.075 0.09 5<br />

9 1 2 2<br />

2.4 <strong>The</strong> Logarithmic Law of the <strong>Wall</strong><br />

<strong>The</strong> wall function approach spares the CFD code from resolving the be-<br />

haviour of the flow in the near-wall region. <strong>The</strong> standard wall function<br />

approach is <strong>to</strong> assume that the the flow behaviour at a certain distance away<br />

from the wall will match the logarithmic law of the wall. <strong>The</strong> governing<br />

equations are solved in the bulk of the flow regime, with the logarithmic law<br />

of the wall providing a boundary condition near the wall.<br />

<strong>Wall</strong> functions are often used with the high-Reynolds-number standard k-ε<br />

model. <strong>The</strong> wall function approach is essentially an alternative <strong>to</strong> the low-<br />

Reynolds-number approach, providing computational savings at the cost of<br />

reduced accuracy. <strong>Wall</strong> functions may be applied <strong>to</strong> the k-ω model, but,<br />

since the advantages of the k-ω model are primarily in its behaviour near the<br />

wall, this is seldom done in practice. <strong>Wall</strong> functions on k, ε, and wall-parallel<br />

velocity, 〈U〉 will be discussed here from a conceptual standpoint. A detailed<br />

discussion of the implementation of log law boundary conditions is left until<br />

Chapter 4.


CHAPTER 2. TURBULENCE MODELS 27<br />

<strong>The</strong> logarithmic law of the wall assumes a logarithmic relationship between<br />

velocity and displacement away from a solid boundary. <strong>The</strong>se quantities are<br />

expressed nondimensionally as U + and y + respectively. (See the discussion<br />

in Chapter 3 on nondimensionalisation.) This relationship is due <strong>to</strong> von<br />

Kármán [28]. <strong>The</strong> relationship is<br />

U + = 1<br />

κ ln Ey +<br />

(2.39)<br />

where κ is von Kármán’s constant. E is a function of wall roughness, and the<br />

smooth-wall value is used in this thesis. Log-law constants appear in Table<br />

2.3. <strong>The</strong> wall roughness constant, E is not <strong>to</strong> be confused with the E term<br />

in the ˜ε transport equation.<br />

Table 2.3: Log-law constants [83]<br />

κ 0.4187<br />

E 9.793<br />

Table 2.4 offers a qualitative appraisal of the behaviour of flow with respect<br />

<strong>to</strong> the nondimensional wall distance, y + . This highlights the qualitatively<br />

different behaviours exhibited by a fluid as a solid boundary is approached.<br />

<strong>The</strong> log law applies approximately where y + > 30. Where y + < 5, flow is<br />

characterised by Prandtl’s [54] hypothesis, that U + = y + very near the wall.<br />

<strong>The</strong> buffer layer is a region (5 < y + < 30) where neither of these assumptions<br />

holds. In specifying a low-Reynolds-number solution, it is important that the<br />

gradients nearest the wall fall within the viscous sublayer (y + < 5). In a high-<br />

Reynolds-number CFD treatment, it is important that the log-law boundary<br />

conditions are applied at a location where y + > 30.<br />

When used as a boundary condition on a high-Reynolds-number k-ε treat-


CHAPTER 2. TURBULENCE MODELS 28<br />

Table 2.4: <strong>Near</strong>-wall flow regimes adapted from [53]<br />

y + < 5 <strong>The</strong> viscous sublayer. <strong>The</strong> flow is essentially laminar.<br />

5 < y + < 30 <strong>The</strong> buffer layer. <strong>The</strong> log law overestimates U + .<br />

y + > 30 <strong>The</strong> log law region. <strong>The</strong> log law is valid.<br />

ment, the log law’s effect on the x momentum equation can be thought of as<br />

a source resulting from shear stress at the wall (τw). <strong>The</strong> relationship can be<br />

shown by considering the definition of U + . (See Chapter 3.)<br />

U + = 〈U〉<br />

Uτ =<br />

Uτ<br />

τw<br />

ρ<br />

(2.40)<br />

(2.41)<br />

Following Launder & Spalding [34], C 1/4<br />

µ k 1/2 may be taken as a velocity scale.<br />

<strong>The</strong> log law estimates the parameter Uτ by<br />

<strong>The</strong>refore,<br />

Uτ = U 2 τ<br />

Uτ<br />

Uτ = C 1/4<br />

µ k 1/2<br />

= τw/ρ 〈U〉<br />

=<br />

1/2 U +<br />

C 1/4<br />

µ k<br />

τw = ρC1/4 µ k1/2 〈U〉<br />

U +<br />

(2.42)<br />

(2.43)<br />

(2.44)<br />

<strong>The</strong> shear stress source in the equation of 〈U〉 is thus known from Equation<br />

2.44, Equation 2.39, and the previous iteration value of 〈U〉 at the point<br />

where the log law is applied. See Chapter 4 for a discussion of the numerical<br />

implementation of the log law.


CHAPTER 2. TURBULENCE MODELS 29<br />

In applying a wall-function boundary condition <strong>to</strong> k, the goal is not simply<br />

<strong>to</strong> prescribe a value of k at the near-wall cell node. Rather, better accuracy<br />

is achieved by applying sources <strong>to</strong> the k equation at the near-wall cell that<br />

represent cell-averaged production and dissipation terms (Pk and ε). Various<br />

methods exist for estimating these quantities. <strong>The</strong> Launder & Spalding [34]<br />

wall function is employed in this work. In this wall function,<br />

<br />

∂ 〈U〉<br />

Pk = τw<br />

∂y<br />

<br />

(k)<br />

ε = ρCµ<br />

2<br />

∂ <br />

〈U〉<br />

∂y<br />

τw<br />

(2.45)<br />

(2.46)<br />

<strong>The</strong> transport equation for ε does not make use of ε. Instead, ε is prescribed<br />

in the near-wall cell from k according <strong>to</strong> the mixing length hypothesis,<br />

C 3/4<br />

3/2 k<br />

µ = κy (2.47)<br />

ε<br />

<strong>The</strong> log law appears in Figure 2.1, plotted <strong>to</strong>gether with DNS and experi-<br />

mental data at various Reynolds numbers. It can be seen that the log law<br />

provides reasonable agreement with DNS and experimental results beyond<br />

the buffer layer (y + > 30), and particularly when y + is not excessively large.<br />

<strong>The</strong> log law produces good results when applied as a boundary condition in<br />

a code whose near-wall cell extends <strong>to</strong> a height that fully encompasses the<br />

buffer layer. Since y + is a function of the flow rate as well as y, the optimum<br />

choice of near-wall cell size may be influenced by the flow rate considered.<br />

However, it is usually sufficient <strong>to</strong> choose a large enough near-wall cell <strong>to</strong><br />

fully encompass the buffer layer for the lowest flow rate anticipated.


CHAPTER 2. TURBULENCE MODELS 30<br />

U +<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

10 0<br />

10 1<br />

+<br />

+<br />

+<br />

+<br />

the log-law<br />

DNS: Reτ = 180 [30]<br />

DNS: Reτ = 395 [44]<br />

DNS: Reτ = 584 [44]<br />

◦ Exp.: Reτ = 708 [87]<br />

▽ Exp.: Reτ = 921 [47]<br />

+<br />

10 2<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

x +<br />

x xxx<br />

x<br />

x<br />

x<br />

x +<br />

x +<br />

+<br />

10 3<br />

y +<br />

10 4<br />

Exp.: Reτ = 1017 [47]<br />

+ Exp.: Reτ = 1655 [47]<br />

⋄ Exp.: Reτ = 2340 [11]<br />

× Exp.: Reτ = 4800 [11]<br />

△ Exp.: Reτ = 8150 [11]<br />

Figure 2.1: <strong>The</strong> log-law compared <strong>to</strong> experiments in channel flow


Chapter 3<br />

<strong>Channel</strong> <strong>Flow</strong><br />

<strong>Channel</strong> flow refers <strong>to</strong> an arrangement where fluid flows between two parallel<br />

walls. Consider walls that are a distance 2δ apart. <strong>Flow</strong> is driven by a<br />

pressure gradient in the x direction and solved from y = 0 at a wall <strong>to</strong> y = δ<br />

at the axis of symmetry between the two walls. <strong>The</strong> flow field is long in<br />

the third dimension, so that the flow is independent of z. <strong>The</strong> flow is also<br />

independent of x. Even in unsteady channel flow, the flow is taken <strong>to</strong> develop<br />

through time but not through the length of the flow direction, x.<br />

3.1 Governing Equations<br />

For ease of explanation, the RANS equations for continuity and conservation<br />

of momentum (Equations 2.4 and 2.5) may be expanded in<strong>to</strong> scalar form.<br />

31


CHAPTER 3. CHANNEL FLOW 32<br />

Removing all velocity gradients in x and z, this yields<br />

Continuity :<br />

X − momentum :<br />

Y − momentum :<br />

∂〈U〉<br />

∂t<br />

∂〈V 〉<br />

∂t<br />

+ 〈V 〉 ∂〈U〉<br />

∂y<br />

+ 〈V 〉 ∂〈V 〉<br />

∂y<br />

∂〈V 〉<br />

∂y<br />

1 ∂〈P 〉 ∂<br />

= − + ρ ∂x ∂y<br />

1 ∂〈P 〉 ∂<br />

= − + ρ ∂y ∂y<br />

= 0 (3.1)<br />

<br />

ν ∂〈U〉<br />

<br />

− ∂y<br />

∂〈uv〉<br />

(3.2)<br />

∂y<br />

<br />

∂〈V 〉<br />

ν − ∂y<br />

∂〈v2 〉<br />

(3.3)<br />

∂y<br />

A boundary condition affecting continuity, Equation 3.1 is that of zero flow<br />

through a solid boundary<br />

〈V 〉| y=0 = 0 (3.4)<br />

<strong>The</strong>refore, Equation 3.1 produces the result that V = 0 throughout the flow<br />

field. <strong>The</strong> momentum equations become<br />

X − momentum :<br />

∂〈U〉<br />

∂t<br />

<br />

1 ∂〈P 〉 ∂<br />

= − + ρ ∂x ∂y<br />

ν ∂〈U〉<br />

∂y<br />

Y − momentum : 0 = − 1 ∂〈P 〉<br />

ρ ∂y − ∂〈v2 〉<br />

∂y<br />

<br />

− ∂〈uv〉<br />

∂y<br />

(3.5)<br />

(3.6)<br />

It is notable, at this stage, that convection has been eliminated from the<br />

momentum equations.<br />

Equation 3.6 can be integrated with respect <strong>to</strong> y and the following boundary<br />

conditions applied<br />

v 2 y=0 = 0 (3.7)<br />

〈P 〉| y=0 = Pw<br />

(3.8)<br />

where Pw is pressure at the wall. Velocities, including turbulent fluctuations,<br />

are zero at the wall, so there are no pressure fluctuations. <strong>The</strong>refore, Pw<br />

is not shown as a mean value. Furthermore, since the flow is driven by a<br />

pressure gradient in x, the pressure at the wall must only vary as a function


CHAPTER 3. CHANNEL FLOW 33<br />

of x and time, hence Pw(x, t). Integrating Equation 3.6 with respect <strong>to</strong> y<br />

and applying the appropriate boundary conditions yields<br />

ρ v 2 + [〈P 〉 − Pw(x, t)] = 0 (3.9)<br />

Equation 3.9 can be differentiated with respect <strong>to</strong> x <strong>to</strong> give<br />

Thus<br />

∂〈P 〉<br />

∂x<br />

∂ 〈P 〉<br />

∂x<br />

is not a function of y.<br />

= ∂<br />

∂x Pw(x, t) (3.10)<br />

In the above analysis, continuity has served <strong>to</strong> simplify the momentum equa-<br />

tions, and y-momentum has yielded the insight that 〈P 〉 varies only in the x<br />

direction and in time. <strong>The</strong> remaining x-momentum equation is solved in the<br />

CFD code. Applying the EVM <strong>to</strong> Equation 3.5, this becomes:<br />

<br />

<br />

∂ 〈U〉 ∂ 〈P 〉 ∂ ∂ 〈U〉<br />

= −1 + (ν + νt)<br />

∂t ρ ∂x ∂y<br />

∂y<br />

(3.11)<br />

In summary, channel flow is governed by a single momentum equation con-<br />

taining terms for fluid acceleration, shear stress, and a driving pressure gra-<br />

dient. Furthermore, convection does not take place in any channel flow, since<br />

wall-normal velocity is zero throughout and all gradients in the wall-parallel<br />

direction are zero.<br />

3.1.1 <strong>The</strong> k-ε Model<br />

In channel flow, the transport equations in the k-ε model (Equations 2.16 &<br />

2.33) become<br />

k :<br />

˜ε : ∂ ˜ε<br />

∂t<br />

<br />

∂k ∂ ν+νt<br />

= ∂t ∂y σk<br />

<br />

∂ ν+νt ∂ ˜ε<br />

˜ε<br />

= + Cε1f1<br />

∂y σε ∂y<br />

k<br />

<br />

∂k<br />

∂y<br />

+ Pk − ε (3.12)<br />

<br />

˜ε 2<br />

Pk − Cε2f2 + E + Y (3.13)<br />

k


CHAPTER 3. CHANNEL FLOW 34<br />

E and Y in Equation 3.13 are unaffected by application <strong>to</strong> channel flow ge-<br />

ometry. (See Equations 2.27 & 2.34.) However, Pk (Equation 2.20) becomes<br />

3.1.2 <strong>The</strong> k-ω Model<br />

Pk = νt<br />

∂ 〈U〉<br />

<strong>The</strong> k-ω transport equations (Equations 2.35 & 2.38) for channel flow are<br />

k :<br />

ω : ∂ω<br />

∂t<br />

∂k<br />

∂t<br />

∂y<br />

<br />

∂ ν+νt ∂k<br />

= ∂y σkω ∂y<br />

<br />

∂ ν+νt ∂ω<br />

= ∂y σω ∂y<br />

with Pk as defined in Equation 3.14.<br />

3.2 <strong>Flow</strong> Characterisation<br />

2<br />

(3.14)<br />

<br />

+ Pk − ωkβ ∗ (3.15)<br />

+ γ <br />

ω<br />

Pk − βω k<br />

2 (3.16)<br />

Steady channel flow is usually characterised by a quantity called the friction<br />

velocity Reynolds number, Reτ. Unsteady channel flow is usually charac-<br />

terised by a time-mean Reτ and an amplitude of oscillation expressed as a<br />

proportion of this time-mean value. To explain Reτ, it is necessary <strong>to</strong> further<br />

analyze steady channel flow.<br />

For steady flow, ∂〈U〉<br />

∂t<br />

= 0 in Equation 3.11 and d〈U〉<br />

dy is not a partial derivative.<br />

In this case, the x-momentum equation (3.5) indicates a balance between the<br />

force acting on the fluid because of the driving pressure gradient and a shear<br />

force resulting from friction. <strong>The</strong> shear stress is a function of velocity, and it


CHAPTER 3. CHANNEL FLOW 35<br />

varies only in y. <strong>The</strong> shear stress may be expressed as<br />

so Equation 3.5 can be rewritten as<br />

d 〈U〉<br />

τ(y) = ρν − ρ 〈uv〉 (3.17)<br />

dy<br />

dτ<br />

dy<br />

It follows from Equation 3.18 that dτ<br />

= d 〈P 〉<br />

dx<br />

(3.18)<br />

dy is constant. Let τw be the shear stress<br />

at the wall, such that τ| y=0 = τw. Also, τ| y=δ = 0 since d〈U〉<br />

dy<br />

<br />

<br />

y=δ<br />

= 0 and<br />

〈uv〉| y=δ = 0. With these boundary conditions on τ, Equation 3.18 may be<br />

integrated <strong>to</strong> yield<br />

d 〈P 〉<br />

dx<br />

= −τw<br />

δ<br />

(3.19)<br />

Reτ may be defined from τw, the geometry of the channel, and the properties<br />

of the fluid according <strong>to</strong><br />

Reτ =<br />

<br />

δ τw<br />

ν ρ<br />

It is of interest, based on Equations 3.19 and 3.20, that only one of<br />

(3.20)<br />

d〈P 〉<br />

, τw, dx<br />

or Reτ must be specified in order <strong>to</strong> completely define the flow. Usually, Reτ<br />

is the quoted parameter.<br />

Another Reynolds number is the bulk Reynolds number, based on average<br />

velocity<br />

Re♭ =<br />

where U indicates a spatial average.<br />

U (2δ)<br />

ν<br />

A further Reynolds number is that based on free stream velocity<br />

Re0 = U0δ<br />

ν<br />

(3.21)<br />

(3.22)


CHAPTER 3. CHANNEL FLOW 36<br />

Re♭ and Re0 are often quoted by experimentalists, who favour these quanti-<br />

ties because they can be readily and directly measured.<br />

3.3 Nondimensionalisation<br />

Viscous velocity and length scales can be defined such that the Reynolds<br />

number based on these scales is equal <strong>to</strong> 1. <strong>The</strong> velocity scale is<br />

and the length scale is<br />

Uτ =<br />

τw<br />

ρ<br />

<br />

ρ<br />

δv = ν<br />

τw<br />

<strong>The</strong> friction velocity Reynolds number may be expressed as<br />

Reτ = Uτδ<br />

ν<br />

= δ<br />

δv<br />

(3.23)<br />

(3.24)<br />

(3.25)<br />

<strong>Flow</strong> variables may be nondimensionalised by the viscous scales in Equations<br />

3.24 and 3.23. By convention, U + , the nondimensional average U velocity is<br />

not shown with braces – 〈〉.<br />

y + = y<br />

δν<br />

<br />

Uτ<br />

= y<br />

ν<br />

U + = 〈U〉<br />

Uτ<br />

〈uv〉 + = 〈uv〉<br />

U 2 τ<br />

k + = k<br />

U 2 τ<br />

(3.26)<br />

(3.27)<br />

(3.28)<br />

(3.29)


CHAPTER 3. CHANNEL FLOW 37<br />

u 2 + = 〈u 2 〉<br />

U 2 τ<br />

v 2 + = 〈v 2 〉<br />

U 2 τ<br />

Sometimes the driving pressure gradient is nondimensionalised by<br />

<br />

P + x =<br />

<br />

∂〈P 〉<br />

∂x<br />

2<br />

(Uτ ) ρ<br />

δ<br />

(3.30)<br />

(3.31)<br />

(3.32)<br />

Note that 〈u 2 〉 and 〈v 2 〉 may be nondimensionalised as 〈u2 〉<br />

k<br />

also notable that the definitions of Reτ and y + give<br />

y + |y=δ = Reτ<br />

3.4 Steady <strong>Channel</strong> <strong>Flow</strong> Data<br />

and 〈v2 〉<br />

. It is<br />

k<br />

(3.33)<br />

Based on the wealth of experimental and DNS results available concerning<br />

the behaviour of steady channel flows, a set of empirical profiles have been<br />

developed <strong>to</strong> obtain estimates of U + , − 〈uv〉 + , k + , 〈u 2 〉 + , and 〈v 2 〉 + . <strong>The</strong>se<br />

results are compared <strong>to</strong> DNS and experimental data in the following sections.<br />

Much of the data were found in the AGARD Aerospace Database 1 . <strong>The</strong><br />

contribu<strong>to</strong>rs of specific data sets are cited in the graph legends 2 .<br />

1 Advisory Group for Aerospace Research and Development, France<br />

2 Another excellent source of experimental information may be found at<br />

http://www.efluids.com. Furthermore, at http://<strong>to</strong>rroja.dmt.vpm.es Dr. Javier Jimenez<br />

(School of Aeronautics, University of Madrid) has published full flow fields for channel<br />

flow at Reτ = 180 and Reτ = 550


CHAPTER 3. CHANNEL FLOW 38<br />

It is intended that these profiles will be used as an aid in developing new<br />

turbulence models or numerical approaches. One particular application is as<br />

a means of setting initial conditions or inlet boundary conditions for a code,<br />

<strong>to</strong> facilitate easier convergence during debugging. A further application is<br />

<strong>to</strong> apply these relations within the code as a means of verifying the internal<br />

calculations of the turbulence model.<br />

<strong>The</strong> profiles shown below are designed <strong>to</strong> match fully turbulent flow and also<br />

laminar/turbulent transitional flow. <strong>The</strong> profiles are reasonably accurate, as<br />

compared <strong>to</strong> the variation generally seen between different sets of experimen-<br />

tal results. It is notable that, in general, flow parameters are strong functions<br />

of Reynolds number. While experimental and DNS results are only available<br />

at a selection of discrete Reynolds numbers, these empirical profiles offer<br />

a means of interpolation and allow codes <strong>to</strong> be validated at intermediate<br />

Reynolds numbers.<br />

One limitation of these emperical profiles is that they do not adequately cap-<br />

ture the limiting shape of turbulence quantities as y + → 0. <strong>The</strong> principal<br />

concern in developing the profiles was <strong>to</strong> accurately account for the magni-<br />

tude and location of peak values. <strong>The</strong>refore, separate near-wall profiles are<br />

also presented.<br />

〈u 2 〉 + and 〈v 2 〉 + are found by multiplying empirical profiles for 〈u2 〉 +<br />

〈v 2 〉 +<br />

k +<br />

k +<br />

and<br />

by the empirical profile for k + . <strong>The</strong> reason why 〈u 2 〉 + and 〈v 2 〉 + are<br />

normalised by k + rather than being fit directly is that this process reduces the<br />

Reynolds number dependence of the resulting empirical profiles, with most<br />

of the Reynolds number dependence in 〈u 2 〉 + and 〈v 2 〉 + being accounted for<br />

in k + . <strong>The</strong> complexity of the empirical profiles is therefore reduced.


CHAPTER 3. CHANNEL FLOW 39<br />

3.4.1 Empirical Profile for U +<br />

Reichardt’s log law [58] states that<br />

U + = 1<br />

κ ln 1.0 + 0.4y +<br />

<br />

+7.8 1 − exp − y+<br />

<br />

+ y<br />

− exp −<br />

11 11<br />

y+<br />

<br />

3<br />

(3.34)<br />

This is plotted against experimental and DNS data in Figure 3.1. Reichardt’s<br />

law follows DNS results very closely near the wall. As y + increases, Re-<br />

ichardt’s law approaches the log law, while experimental values of U + tend<br />

<strong>to</strong> fall above the log law for high y + .<br />

3.4.2 Empirical Profile for − 〈uv〉 +<br />

Based on Reichardt’s law, a profile can be derived for − 〈uv〉 + . This was<br />

originally done by Alexander Davroux and Dominique Laurence at Electricité<br />

de France.<br />

dτ<br />

dy is a constant, τ| y=0 = τw and τ| y=δ = 0. <strong>The</strong>refore, the equation for τ (y)<br />

in a channel is<br />

τ (y) = τw<br />

<br />

1 − y<br />

<br />

δ<br />

This can be inserted in<strong>to</strong> Equation 3.17 <strong>to</strong> give<br />

− 〈uv〉 = τw<br />

ρ<br />

In nondimensional form, this becomes<br />

<br />

1 − y<br />

<br />

δ<br />

− 〈uv〉 + <br />

= 1 − y<br />

<br />

δ<br />

−<br />

− ν dU<br />

dy<br />

dU +<br />

dy +<br />

(3.35)<br />

(3.36)<br />

(3.37)


CHAPTER 3. CHANNEL FLOW 40<br />

Substituting Equation 3.34 in<strong>to</strong> Equation 3.37 gives<br />

− 〈uv〉 + =<br />

<br />

1 − y<br />

<br />

0.976<br />

−<br />

δ 1 + 0.4y +<br />

<br />

<br />

+ 0.709 exp − y+<br />

<br />

11<br />

<br />

−0.709 exp − y+<br />

<br />

+ 0.234y<br />

3<br />

+ <br />

∗ exp − y+<br />

<br />

3<br />

in<strong>to</strong> Equation 3.38 gives<br />

− 〈uv〉 + = 1 − y+<br />

<br />

0.976<br />

−<br />

Reτ 1 + 0.4y +<br />

<br />

<br />

− 0.709 exp − y+<br />

<br />

11<br />

+ 0.709 − 0.234y + <br />

exp − y+<br />

<br />

3<br />

Substituting y+<br />

Reτ<br />

for y<br />

δ<br />

(3.38)<br />

(3.39)<br />

Figure 3.2 shows the performance of Reichardt’s law when used <strong>to</strong> calcu-<br />

late − 〈uv〉 + by the above technique. One limitation of this profile is that<br />

− 〈uv〉 + < 0 for low values of y + , particularly when Reτ is small. To im-<br />

prove upon this, the (0.709 − 0.234y + ) component of Equation 3.39 can be<br />

modified <strong>to</strong> alter the behaviour at low y + <br />

. Furthermore, can be<br />

0.976<br />

1+0.4y +<br />

adjusted so that the profile produces a value of zero when y + = 0. <strong>The</strong> result<br />

of these manipulations is<br />

− 〈uv〉 + = 1 − y+<br />

<br />

1<br />

−<br />

Reτ 1 + 0.4y +<br />

<br />

+ 0.709 − 0.18y + <br />

exp − y+<br />

<br />

3<br />

<br />

− 0.709 exp<br />

Figure 3.3 plots Equation 3.40 at various values of Reτ.<br />

3.4.3 Empirical Profile for k +<br />

− y+<br />

11<br />

<br />

(3.40)<br />

<strong>The</strong> following profile for k + is adapted from a profile originally fitted by<br />

Alexander Davroux and Dominique Laurence at Electricité de France. <strong>The</strong>


CHAPTER 3. CHANNEL FLOW 41<br />

profile is<br />

k + =<br />

1<br />

Reτ 2<br />

0.07 + 0.05<br />

∗<br />

1600<br />

y + <br />

2<br />

∗ exp − y+<br />

<br />

7<br />

<br />

+4.5 1 − exp − y+<br />

+ −1 <br />

4y<br />

+ 1<br />

20 Reτ<br />

<br />

+ 2 <br />

y<br />

∗ 1 − exp −<br />

3<br />

Figure 3.4 shows this profile for k + plotted against DNS results.<br />

(3.41)<br />

Table 3.1 offers qualitative descriptions of the contributions of the various<br />

terms in Equation 3.41. <strong>The</strong> final term ensures that k + = 0 at the wall, but<br />

the shape of the profile in the very near-wall region does not closely match<br />

DNS data. An analysis of the near-wall behaviour of turbulence quantities<br />

will follow.<br />

Table 3.1: Qualitative behaviour of terms in Equation 3.41<br />

<br />

Reτ<br />

1600<br />

<br />

0.07 + 0.05<br />

<br />

4.5 1 − exp<br />

<br />

∗ (y + ) 2 <br />

∗ exp − y+<br />

<br />

7<br />

<br />

4y + −1<br />

+ 1<br />

− y+<br />

20<br />

<br />

1 − exp −<br />

y +<br />

3<br />

Reτ<br />

2 <br />

3.4.4 Empirical Profile for u 2 +<br />

<strong>The</strong> following formula for 〈u2 〉 +<br />

k +<br />

fits the peak value of k +<br />

helps <strong>to</strong> roughly capture the curve<br />

at high Reynolds numbers<br />

forces a value of zero at the wall<br />

was adapted from a profile fitted by Alexan-<br />

der Davroux and Dominique Laurence at Electricité de France.


CHAPTER 3. CHANNEL FLOW 42<br />

〈u 2 〉 +<br />

k +<br />

+ 1/4<br />

y<br />

= 1.46 − 0.542 ∗ +<br />

Reτ<br />

⎛<br />

y +<br />

⎜<br />

⎝<br />

1.5 +<br />

<br />

7<br />

2 y +<br />

7<br />

⎞<br />

⎟<br />

⎠ (3.42)<br />

Figure 3.5 shows the profile for 〈u 2 〉 + obtained by multiplying Equation 3.42<br />

by Equation 3.41. This is plotted against DNS results.<br />

Table 3.2 offers qualitative descriptions of the contributions of the various<br />

terms in Equation 3.42.<br />

Table 3.2: Qualitative behaviour of terms in Equation 3.42<br />

„ y +<br />

«<br />

<br />

7<br />

“ ”<br />

y + 2<br />

1.5+ 7<br />

<br />

y + 1/4<br />

−0.542 ∗<br />

Reτ<br />

fits the peak value of 〈u2 〉 +<br />

k +<br />

shifts the curve at higher y + <strong>to</strong> account for some<br />

Reynolds number dependence of 〈u2 〉 +<br />

3.4.5 Empirical Profile for v 2 +<br />

This profile for 〈v2 〉 +<br />

k +<br />

was originally fitted by Alexander Davroux and Do-<br />

minique Laurence at Electricité de France. <strong>The</strong> profile is<br />

〈v 2 〉 +<br />

k +<br />

where<br />

=<br />

<br />

y + 2<br />

<br />

<br />

<br />

+<br />

20<br />

y<br />

0.08 ∗ 2 + 0.8 ∗ 1λR + λR ∗ 0.06 ∗ + 1<br />

y +<br />

Reτ<br />

2 + 20<br />

<br />

+ 2 <br />

+ 2<br />

−1<br />

y y<br />

∗ 1 − exp −<br />

∗ 1 − exp −<br />

(3.43)<br />

20<br />

3<br />

<br />

λR = exp − 100<br />

<br />

Reτ<br />

k +


CHAPTER 3. CHANNEL FLOW 43<br />

Figure 3.6 shows the resulting profile for 〈v 2 〉 + plotted against DNS results.<br />

3.4.6 <strong>Near</strong>-<strong>Wall</strong> Behaviour<br />

Various researchers have analysed the limiting behaviour of turbulence pa-<br />

rameters near a wall [4, 8, 22, 39, 43, 81]. Following Hanjalic & Launder [22],<br />

fluctuating velocity components can be represented in the following way as<br />

y → 0.<br />

u + rms = a ∗ y + y +<br />

+ O<br />

2 <br />

v + rms = b ∗ y + <br />

2 y +<br />

+ O<br />

3 <br />

(3.44)<br />

where a and b are constants, and u + rms and v + rms are root-mean-square aver-<br />

aged values.<br />

<strong>The</strong>se relationships are supported by the DNS results of Kim et. al. [30].<br />

As Suga [77] points out:<br />

2<br />

u + 2<br />

= a ∗ y + <br />

2 y +<br />

+ O<br />

4 <br />

2<br />

v + 2<br />

= b ∗ y + <br />

4 y +<br />

+ O<br />

6 <br />

〈uv〉 + = 〈ab〉 ∗ y + <br />

3 y +<br />

+ O<br />

4 <br />

(3.45)<br />

Thus the near-wall behaviour of 〈u 2 〉 is expected <strong>to</strong> be dominated by a curve


CHAPTER 3. CHANNEL FLOW 44<br />

of the form (y + ) 2 ; 〈v 2 〉 by (y + ) 4 ; -〈uv〉 by (y + ) 3 . Specifically, as y + → 0,<br />

where,<br />

〈u2 〉 +<br />

(y + 2<br />

)<br />

→ A<br />

〈v2 〉 +<br />

(y + 4<br />

)<br />

→ B<br />

− 〈uv〉 +<br />

(y + 3<br />

)<br />

→ C (3.46)<br />

A = a 2<br />

B = b 2<br />

C = − 〈ab〉 (3.47)<br />

Based on the definition of k + , it can be shown that [22]<br />

where γ is a constant in,<br />

k + = 1 2<br />

a<br />

2<br />

+ γ 2 ∗ y + <br />

2 y +<br />

+ O<br />

3 <br />

2<br />

w = γ 2 ∗ y + <br />

2 y +<br />

+ O<br />

4 <br />

Thus, k + is dominated by (y + ) 2 near the wall and, as y + → 0,<br />

where,<br />

(3.48)<br />

(3.49)<br />

k +<br />

(y + 2 → D (3.50)<br />

)<br />

D = 1 2<br />

a<br />

2<br />

+ γ 2<br />

Figure 3.7 shows the near-wall behaviours of 〈u2 〉 +<br />

(3.51)<br />

(y + ) 2 , 〈v2 〉 +<br />

(y + ) 4 , −〈uv〉+<br />

(y + ) 3 , and k+<br />

(y + ) 2 ,<br />

respectively. <strong>The</strong>se graphs are expected <strong>to</strong> tend <strong>to</strong> horizontal lines as y + → 0.


CHAPTER 3. CHANNEL FLOW 45<br />

<strong>The</strong> profiles quoted above for 〈u 2 〉 + , 〈v 2 〉 + , − 〈uv〉 + , and k + near the wall do<br />

not adequately satisfy the relationships in 3.46 and 3.50. <strong>The</strong>refore, in the<br />

interest of completeness, values of A, B, C, and D are found as functions of<br />

Reynolds number based on DNS results 3 . [30, 44]<br />

<strong>The</strong> functions are:<br />

A = 1.060 × 10 −4 Reτ + (0.1107)<br />

B = 1.757 × 10 −7 Reτ + 4.546 × 10 −5<br />

C = 9.005 × 10 −7 Reτ + 5.855 × 10 −4<br />

D = 9.362 × 10 −5 Reτ + (0.0680) (3.52)<br />

<strong>The</strong> Reynolds-number dependent near-wall proportionalities indicated above<br />

are consistent with the DNS results of An<strong>to</strong>nia & Kim [4], who analyzed<br />

the near-wall behaviour of u + , v + , − 〈uv〉 + , and other parameters at two<br />

Reynolds numbers.<br />

3.5 Local Nondimensionalisation<br />

When examining computed results, it is often more straightforward <strong>to</strong> nor-<br />

malise y by k, rather than by τw. In a complex, multidimensional mesh, it is<br />

not possible, in general, <strong>to</strong> relate modelled flow parameters <strong>to</strong> friction at a<br />

corresponding point along the wall.<br />

When y is normalised by k, the cus<strong>to</strong>mary notation is y ∗ . <strong>The</strong> method of<br />

3 It is notable that the relationships A = a 2 , B = b 2 , and C = 〈ab〉 do not imply<br />

C = √ AB.


CHAPTER 3. CHANNEL FLOW 46<br />

obtaining y ∗ is<br />

y ∗ can be related <strong>to</strong> y + by<br />

y ∗ = y√ k<br />

ν<br />

Likewise, 〈U〉 can be normalised by k.<br />

(3.53)<br />

y ∗ = y +√ k + (3.54)<br />

U ∗ = 〈U〉<br />

√ k<br />

U ∗ =<br />

U +<br />

√ k +<br />

(3.55)<br />

(3.56)<br />

y + is plotted against y ∗ for various values of Reτ in Figure 3.8. U + and U ∗ at<br />

various Reynold’s numbers are plotted in Figure 3.9. It can be seen that U +<br />

and U ∗ can differ significantly. This may be caused by the fact that k + is not<br />

increasing mono<strong>to</strong>nically. An interesting feature, visible in Figure 3.9, is that<br />

U ∗ approaches a constant as y → 0. This happens because, as y → 0, U +<br />

becomes proportional <strong>to</strong> y + and k + becomes proportional <strong>to</strong> (y + ) 2 . Thus,<br />

U +<br />

√ k + = U ∗ approaches a constant.<br />

Another choice of variable against which <strong>to</strong> normalise y and U is 〈v 2 〉. 〈v 2 〉<br />

has the same units as k. Since most of the local variability of k results from<br />

turbulent fluctuations in the mean flow direction (〈u 2 〉), local nondimension-<br />

alisation based on 〈v 2 〉 may be expected <strong>to</strong> offer smoother profiles. We can<br />

define<br />

y ∗<br />

v 2 = y 〈v 2 〉<br />

ν<br />

U ∗ 〈U〉<br />

v2 = =<br />

〈v2 〉<br />

= y +<br />

<br />

〈v2 〉 +<br />

U +<br />

<br />

〈v 2 〉 +<br />

(3.57)<br />

(3.58)


CHAPTER 3. CHANNEL FLOW 47<br />

y + is plotted against y ∗<br />

v 2 for various values of Reτ in Figure 3.10. <strong>The</strong>re is<br />

reasonable agreement between y + and y ∗<br />

v 2. U + and U ∗<br />

v 2 at various Reynold’s<br />

numbers are plotted in Figure 3.11. <strong>The</strong> 〈v 2 〉 normalising approach suffers<br />

from the fact that 〈v 2 〉 + tends <strong>to</strong> zero as y → 0 faster than U + does, so that<br />

U ∗<br />

v 2 becomes large. As y → 0, U + is expected <strong>to</strong> become proportional <strong>to</strong> y +<br />

and 〈v2 〉 + becomes proportional <strong>to</strong> (y + ) 4 U , so +<br />

√<br />

〈v2 +<br />

〉<br />

1<br />

y + .<br />

becomes proportional <strong>to</strong>


Chapter 4<br />

Numerical Implementation<br />

This work uses a code derived from PASSABLE. 1 Prior <strong>to</strong> this work, the<br />

code was modified for periodic flow by Addad [1]. A finite volume method<br />

is used <strong>to</strong> discretise the equations governing the flow. More information on<br />

the numerical implementation of CFD can be found in books. [83],[20]<br />

4.1 <strong>The</strong> Mesh<br />

As discussed in Chapter 3, the flow field is a one-dimensional expanse with<br />

a wall at y = 0 and symmetry at y = δ. This is modelled as a channel ’slice’<br />

of finite thickness. <strong>The</strong> slice is divided in<strong>to</strong> finite volumes, called ‘cells’.<br />

<strong>The</strong> slice is one cell thick in x and is divided in y according <strong>to</strong> the degree<br />

of numerical accuracy required. This is shown schematically in Figure 4.1.<br />

<strong>The</strong> actual grid used for low-Reynolds-number solutions within this work<br />

1 PArabolic Solution Scheme <strong>Applied</strong> <strong>to</strong> Boundary Layer Equations<br />

48


CHAPTER 4. NUMERICAL IMPLEMENTATION 49<br />

contains 98 cells. Each cell is 5% larger than its wall-facing neighbour. This<br />

geometrical feature is referred <strong>to</strong> as an expansion fac<strong>to</strong>r of 1.05.<br />

†<br />

d<br />

y<br />

U P<br />

symmetry<br />

nodes<br />

vertexes<br />

wall<br />

Figure 4.1: <strong>The</strong> low-Reynolds-number grid<br />

Within each cell is a ‘node’, where parameters associated with the cell are<br />

tracked. Cell boundaries, called ‘vertexes’, are centred between nodes. Ex-<br />

ceptions exist at y = 0 and y = δ. At y = δ, a half-cell exists, with the<br />

final node placed at the symmetry plane. At y = 0, a node is placed at<br />

the wall with no associated cell. This represents a new modification <strong>to</strong> the<br />

PASSABLE code, allowing easier implementation of wall function boundary<br />

conditions. Previously, the code employed a similar mesh style at y = 0 <strong>to</strong><br />

the mesh at y = δ.<br />

Table 4.1 shows the notation used in this thesis <strong>to</strong> refer <strong>to</strong> discretised values.<br />

Values at vertexes are not s<strong>to</strong>red in the code, but may be calculated via<br />

linear interpolation from adjacent nodal values.


CHAPTER 4. NUMERICAL IMPLEMENTATION 50<br />

Table 4.1: <strong>The</strong> notation for discretised values<br />

current node subscript P<br />

node above subscript N<br />

node below subscript S<br />

vertex above subscript n<br />

vertex below subscript s<br />

previous time step superscript t − 1<br />

cell height ∆yp<br />

node-<strong>to</strong>-node dist. ∆yn,s<br />

time step ∆t<br />

4.2 Volume Integral Form<br />

<strong>The</strong> finite volume discretisation method is convenient because quantities are<br />

conserved within each finite volume and are therefore necessarily conserved<br />

throughout the flow field. In order <strong>to</strong> employ the finite volume method, it is<br />

first necessary <strong>to</strong> integrate the differential equations governing the flow with<br />

respect <strong>to</strong> volume. <strong>The</strong> volume-integrated forms of the equations are pre-<br />

sented below. Note that, in channel flow geometry, integrating with respect<br />

<strong>to</strong> volume is equivalent <strong>to</strong> integrating with respect <strong>to</strong> y.<br />

〈U〉 :<br />

k :<br />

˜ε :<br />

∂〈U〉<br />

∂k<br />

∂t<br />

∂〈P 〉<br />

∂〈U〉<br />

dy + (ν + νt) ∂x ∂y<br />

<br />

ν+νt ∂k<br />

∂t dy = − 1<br />

ρ<br />

dy =<br />

<br />

∂ ˜ε dy = ∂t<br />

ω : ∂ω dy = ∂t<br />

− Cε2f2<br />

ν+νt<br />

σk<br />

<br />

ν+νt ∂ ˜ε<br />

σω<br />

σε<br />

<br />

˜ε 2<br />

(4.1)<br />

∂y + Pkdy − εdy (4.2)<br />

∂y + <br />

˜ε<br />

Cε1f1 Pkdy k<br />

<br />

dy + k<br />

Edy + Y dy (4.3)<br />

<br />

∂ω<br />

∂y + γ <br />

ω Pkdy − k<br />

βω2dy (4.4)


CHAPTER 4. NUMERICAL IMPLEMENTATION 51<br />

Before performing any dimensional analysis on the above equations, one must<br />

actually integrate with respect <strong>to</strong> volume, applying limits of 0 <strong>to</strong> 1 on the in-<br />

tegrations in x and z. It is also notable that the PASSABLE code uses a grid<br />

that is unit length in z but not unit length in x (∆x = 1). This detail must<br />

be remembered when working with the code, particularly when specifying<br />

volume-integrated source terms. Furthermore, in the PASSABLE implemen-<br />

tation, Equations 4.1-4.4 are multiplied through by ρ. This convention is<br />

maintained in the subgrid.<br />

4.3 Discretisation<br />

<strong>The</strong> process of discretisation involves approximating the governing differen-<br />

tial equations by algebraic equations so that they can be solved by a com-<br />

puter.<br />

A fully implicit scheme is used in this work. This means that the previous<br />

time-step values of adjacent nodes are not used. In contrast, a fully explicit<br />

scheme would use the previous time-step values, but not the current time-<br />

step values of adjacent nodes. <strong>The</strong> fully implicit method ensures that the<br />

solution can converge at large time steps.<br />

Central differencing is used in this work <strong>to</strong> approximate derivatives. Other<br />

differencing schemes include PLDS 2 [51] and QUICK 3 [37]. <strong>The</strong>se alternative<br />

schemes involve higher-order approximations. <strong>The</strong> QUICK scheme is also<br />

upstream-biased, in the sense that it makes use of more information on the<br />

2 Power-Law Differencing Scheme<br />

3 Quadratic Upstream Interpolation for Convective Kinetics


CHAPTER 4. NUMERICAL IMPLEMENTATION 52<br />

upstream side of the derivative than on the downstream side. (<strong>The</strong> scheme is<br />

influenced by the local direction of 〈U〉.) This is physically reasonable, since<br />

quantities are convected downstream. Because of the simple geometry and<br />

the lack of convective terms in the governing equations, central differencing<br />

was felt <strong>to</strong> be an effective choice in this work.<br />

Within the CFD code, a system of algebraic equations is compiled <strong>to</strong> repre-<br />

sent each of the momentum equations or transport equations <strong>to</strong> be solved.<br />

A set of algebraic equations representing one differential equation is solved<br />

using an iterative procedure 4 before the next differential equation is treated.<br />

<strong>The</strong> various differential equations are coupled <strong>to</strong> one-another, so the entire<br />

process of solving each differential equation in sequence must be repeated<br />

iteratively until the overall solution converges. 5<br />

<strong>The</strong> discretised momentum equation for 〈U〉 is<br />

where<br />

〈U〉 P − 〈U〉 t−1<br />

P<br />

∆t<br />

∆yp = − 1 ∂ 〈P 〉<br />

ρ<br />

<br />

∂ 〈U〉<br />

∂y<br />

<br />

∂ 〈U〉<br />

∂x ∆yp + (ν + νt) n<br />

<br />

∂ 〈U〉<br />

− (ν + νt) s<br />

n<br />

∂y<br />

= 〈U〉 N − 〈U〉 P<br />

∆yn<br />

s<br />

<br />

∂ 〈U〉<br />

∂y<br />

n<br />

(4.5)<br />

(4.6)<br />

∂y<br />

=<br />

s<br />

〈U〉 P − 〈U〉 S<br />

∆ys<br />

(4.7)<br />

and (ν + νt) n and (ν + νt) s are found using linear interpolation between<br />

nodes.<br />

4 <strong>The</strong> iterative solution procedure used for the solution of algebraic equations in this<br />

study is the Tri-Diagonal Matrix Algorithm (TDMA). This procedure is appropriate <strong>to</strong><br />

parabolic systems of equations.<br />

5 An iterative solution algorithm is said <strong>to</strong> be converged when the solution that it<br />

produces is observed <strong>to</strong> undergo no significant change with successive iterations.


CHAPTER 4. NUMERICAL IMPLEMENTATION 53<br />

<strong>The</strong> discretised transport equation for k is<br />

kP − k t−1<br />

P<br />

∆t ∆yp =<br />

<br />

ν + νt<br />

<br />

∂k<br />

∂y<br />

where<br />

σk<br />

n<br />

+Pk∆yp − ε∆yp<br />

Pk = νt<br />

n<br />

〈U〉n − 〈U〉 s<br />

∆yp<br />

and 〈U〉 n and 〈U〉 s are interpolated values.<br />

<br />

ν + νt<br />

−<br />

2<br />

σk<br />

s<br />

<br />

∂k<br />

∂y s<br />

(4.8)<br />

(4.9)<br />

Discretised transport equations for ˜ε and ω can be found in a similar manner.<br />

Some further equations whose discretised forms are noteworthy are<br />

⎛<br />

⎞2<br />

∂〈U〉 ∂〈U〉<br />

− ∂y<br />

∂y<br />

E = 2ννt ⎝ n<br />

s ⎠<br />

∆yp<br />

⎛<br />

ˆε = 2ν ⎝<br />

√k n<br />

−<br />

∆yp<br />

√k s<br />

⎞<br />

⎠<br />

2<br />

(4.10)<br />

(4.11)<br />

Thus, an algebraic equation may be generated for each node P . <strong>The</strong> discre-<br />

tised differential equations are expressed in the code in the form<br />

(AP − SP ) φp = ANφN + ASφS + SU<br />

(4.12)<br />

where φ is the unknown parameter from the original differential equation.<br />

AP , AN and AS are coefficients on nodal values. Previous time step infor-<br />

mation is included as a source. <strong>The</strong> source is split in<strong>to</strong> two terms, SU and<br />

SP φp for reasons of numerical stability. It is advantageous <strong>to</strong> have a large<br />

coefficient on φp, so negative quantities are sometimes moved from SU in<strong>to</strong><br />

SP (dividing by φp) <strong>to</strong> artificially increase this coefficient. At other times,<br />

sources involve a product of φp, and the use of SP is a natural choice.


CHAPTER 4. NUMERICAL IMPLEMENTATION 54<br />

<strong>The</strong> fully implicit scheme also requires information on φ t−1<br />

p . This does not<br />

appear explicitly in Equation 4.12 because it is included in the source term,<br />

SU.<br />

As an example of the expressions of the coefficients presented in Equation<br />

4.12, the coefficients on 〈U〉 are<br />

AN = (ν + νt) n<br />

∆yn<br />

AS = (ν + νt) s<br />

∆ys<br />

AP = (ν + νt) n<br />

∆yn<br />

= AN + AS<br />

SP = − 1<br />

∆t ∆yp<br />

SU =<br />

<br />

〈U〉 t−1<br />

P<br />

∆t<br />

<br />

+ (ν + νt) s<br />

∆ys<br />

∆yp −<br />

4.4 Boundary Conditions<br />

1<br />

ρ<br />

<br />

∂ 〈P 〉<br />

∆yp<br />

∂x<br />

(4.13)<br />

At y = δ, a symmetry boundary condition is employed. <strong>The</strong> governing<br />

equations are not solved at y = δ. Values of 〈U〉, k, ˜ε, and ω are copied <strong>to</strong><br />

the symmetry plane node from the nearest adjacent node. At this adjacent<br />

node, the coefficients are adjusted as follows for each of 〈U〉, k, ˜ε, and ω:<br />

AN = 0<br />

AP = AS (4.14)<br />

This ensures that all gradients in y are zero at the symmetry plane.<br />

At the wall, velocities are zero. This creates large gradients in the near-


CHAPTER 4. NUMERICAL IMPLEMENTATION 55<br />

wall region. Because CFD uses discretisation, the s<strong>to</strong>rage and computational<br />

requirements associated with accurately representing large gradients are high.<br />

This results in a great potential for computation-saving approaches at the<br />

wall. Many methods exist for applying wall boundary conditions <strong>to</strong> CFD<br />

codes, and these may be model-specific. Boundary conditions on the wall<br />

are discussed below.<br />

4.4.1 <strong>Wall</strong> Boundaries on k-ε<br />

In the low-reynolds-number standard k-ε model, ˜ε is defined as being equal<br />

<strong>to</strong> zero at the wall. Furthermore, k = 0 at the wall, arising from the fact that<br />

k represents turbulent kinetic energy and all velocities, including turbulent<br />

fluctuations, are zero at the wall. In the low-Reynolds-number model, these<br />

quantities are simply prescribed <strong>to</strong> be zero at the wall. Transport equations<br />

are not solved at the wall, but are solved in the wall-adjacent cell. Because<br />

of the large gradients involved, finer grid resolution is required near the wall,<br />

as shown in Figure 4.1.<br />

4.4.2 <strong>Wall</strong> Boundaries on k-ω<br />

As with the k-ε model, 〈U〉 = 0 and k = 0 at the wall, and transport<br />

equations may be solved up <strong>to</strong> the wall-adjacent cell. However, ω → ∞ as<br />

y → 0, so it is not numerically possible <strong>to</strong> specify ω at the wall. Wilcox [89]<br />

notes that the limiting behaviour of ω at the wall is<br />

ω → 6ν<br />

βy 2 as y → 0 (4.15)


CHAPTER 4. NUMERICAL IMPLEMENTATION 56<br />

Thus, while ω is infinite at the wall, finite values may be specified near the<br />

wall according <strong>to</strong> Equation 4.15. However, the gradient of ω near the wall<br />

is excessively large (approximately two orders of magnitude greater than<br />

gradients on ˜ε), so it is not practical <strong>to</strong> solve the ω transport equation in the<br />

near wall region. <strong>The</strong> approach recommended by Wilcox [91] is <strong>to</strong> specify<br />

ω according <strong>to</strong> the limiting behaviour suggested in Equation 4.15 within the<br />

first 7 <strong>to</strong> 10 cells, where y + < 2.5. Transport equations on 〈U〉 and k use<br />

specified values of ω in this region, and the transport equation of ω is only<br />

solved outside of this very near-wall region.<br />

4.4.3 <strong>The</strong> Logarithmic Law of the <strong>Wall</strong><br />

A wall function mesh incorporates a near-wall cell that is large enough <strong>to</strong><br />

fully encompass the buffer layer and extend in<strong>to</strong> the region where the log<br />

law is valid. This is shown schematically in Figure 4.2. Because flows of low<br />

Reynolds number are considered in this thesis, the buffer layer is very large.<br />

<strong>The</strong> near-wall node location was set <strong>to</strong> y/δ = 20% or y + = 36.0. <strong>The</strong> main<br />

grid of PASSABLE uses vertexes centred between nodes, so the near-wall<br />

node is positioned far from the centre of its cell. <strong>The</strong> remaining grid was<br />

made up of 18 nodes. An expansion fac<strong>to</strong>r of 1.00 was used for these 18 cells.<br />

<strong>The</strong> <strong>to</strong>tal number of grid cells used with wall functions was thus much less<br />

than with the low-Reynolds-number solution.<br />

Numerically, the log law is implemented by severing the link <strong>to</strong> the wall<br />

in the near-wall cell, and then manipulating the 〈U〉, k, ε source terms in<br />

the near-wall cell <strong>to</strong> reflect the log law. A full discussion of the numerical<br />

implementation of log laws in CFD can be found in the book of Versteeg &


CHAPTER 4. NUMERICAL IMPLEMENTATION 57<br />

Malalasekera [83].<br />

U +<br />

P<br />

nience:<br />

wall<br />

function<br />

applied<br />

Figure 4.2: <strong>The</strong> high-Reynolds-number grid<br />

is calculated from the log law (Equation 2.39), repeated here for conve-<br />

U +<br />

P<br />

1<br />

=<br />

κ ln Ey + <br />

p<br />

(4.16)<br />

where y + p is obtained from the C 1/4<br />

µ k 1/2 velocity scale suggested by Launder<br />

& Spalding [34]:<br />

y + p = C1/4 µ k 1/2<br />

p yp<br />

ν<br />

yp is the distance of the near-wall cell node from the wall.<br />

<strong>Wall</strong> shear stress is obtained from Equation 2.44. It is<br />

− τw<br />

ρ<br />

τw = ρC1/4 µ k 1/2<br />

p 〈U〉 P<br />

U +<br />

P<br />

(4.17)<br />

(4.18)<br />

from Equation 4.18 may be added <strong>to</strong> the x-momentum equation as a


CHAPTER 4. NUMERICAL IMPLEMENTATION 58<br />

source. 6 Since τw contains a 〈U〉 P term, the SP source is used:<br />

SP = S ′ P − C1/4 µ k 1/2<br />

p<br />

U + p<br />

(4.19)<br />

where S ′ P is the value of SP before the log-law term is added (containing only<br />

the time-dependence term).<br />

Also, AS is modified <strong>to</strong> cut the link <strong>to</strong> the wall and prevent a second (in-<br />

accurate) accounting of τw via the normal functioning of the x-momentum<br />

equation. <strong>The</strong> modified terms are<br />

AS = 0<br />

AP = AN (4.20)<br />

This modification is common <strong>to</strong> all wall function treatments on 〈U〉, k, ε,<br />

and ω.<br />

Average production and dissipation of k in the near-wall cell ( <br />

Pk p and<br />

(ε) p ) are calculated from Equations 2.45 and 2.46. In descretised form, these<br />

become<br />

Substituting Equation 4.18,<br />

<br />

〈U〉P<br />

Pk = τw<br />

p<br />

yp<br />

<br />

(kp)<br />

(ε) p = ρCµ<br />

2<br />

〈U〉P <br />

<br />

Pk p<br />

τw<br />

yp<br />

ρC1/4 µ k<br />

= 1/2<br />

p (〈U〉 P ) 2<br />

U +<br />

P<br />

(ε) p = C 3/4<br />

µ k 3/2<br />

· yp<br />

U +<br />

P<br />

yp<br />

<br />

(4.21)<br />

(4.22)<br />

(4.23)<br />

(4.24)<br />

6 In general, one would integrate τw with respect <strong>to</strong> the surface area over which it acts<br />

before applying the result as a source <strong>to</strong> the x-momentum equation. In channel flow<br />

geometry, τw acts over a surface of unit area.


CHAPTER 4. NUMERICAL IMPLEMENTATION 59<br />

<br />

1<br />

1<br />

Pk dy − (ε)p dy<br />

ρ p ρ<br />

must be added <strong>to</strong> the transport equation of k<br />

in the near-wall cell as a source. k may be fac<strong>to</strong>red out of the equation for<br />

(ε) p , and the remaining equation (including the k 1/2 term) may be placed in<br />

the SP source. This amounts <strong>to</strong> modelling k 3/2<br />

p in the equation for (ε) p as<br />

<br />

1/2<br />

kp · k old<br />

p<br />

, where k old<br />

p is the previous iteration value of kp.<br />

<strong>The</strong> terms in the k equation are modified as follows:<br />

SP = S ′ <br />

C<br />

P −<br />

3/4<br />

µ k 1/2<br />

U + <br />

p<br />

P<br />

∆yp<br />

ρ yp<br />

<br />

C 1/4<br />

µ k 1/2<br />

p (〈U〉 P ) 2<br />

<br />

SU = S ′ U +<br />

U +<br />

P<br />

· yp<br />

∆yp<br />

(4.25)<br />

In treating the transport equation of ε, the value of ε is simply prescribed in<br />

the near-wall cell according <strong>to</strong> the mixing length hypothesis (Equation 2.47):<br />

εp = C3/4 µ k3/2 κyp<br />

(4.26)<br />

By making SP and SU large, other terms in the transport equation of ε may<br />

be made numerically insignificant in the near-wall cell, allowing a prescribed<br />

value <strong>to</strong> be specified in SU. <strong>The</strong> source terms are set as follows:<br />

SP = −G<br />

<br />

C<br />

SU = G ·<br />

3/4<br />

µ k3/2 <br />

κyp<br />

(4.27)<br />

where G is a very large number, chosen <strong>to</strong> be several orders of magnitude<br />

greater than any other terms that will appear in the transport equation of ε.


CHAPTER 4. NUMERICAL IMPLEMENTATION 60<br />

4.4.4 <strong>The</strong> Subgrid Approach<br />

<strong>The</strong> subgrid implementation employed in this work begins with the same grid<br />

used for a log-law wall function, discussed above. This is referred <strong>to</strong> as the<br />

‘main grid’. Within the wall-adjacent cell of the main grid, the subgrid code<br />

generates a subgrid mesh, as shown schematically in Figure 4.3. Within<br />

the subgrid mesh, 50 nodes were used, with an expansion fac<strong>to</strong>r of 1.10.<br />

<strong>The</strong> subgrid mesh is similar <strong>to</strong> the low-Reynolds-number mesh shown in<br />

Figure 4.1, but it fills only the near-wall region corresponding <strong>to</strong> the wall-<br />

adjacent cell of the main grid. <strong>The</strong> subgrid mesh differs from the main<br />

grid mesh in the choice of node locations. While the main grid uses cell<br />

vertexes centred between nodes, the subgrid coded in this work uses nodes<br />

centred between vertexes. 7 This affects way in which linear interpolations<br />

are calculated within the code.<br />

main grid<br />

node<br />

subgrid<br />

node<br />

subgrid<br />

region<br />

Figure 4.3: <strong>The</strong> subgrid mesh, adapted from Gant [21]<br />

7 This is due only <strong>to</strong> programmer preference.


CHAPTER 4. NUMERICAL IMPLEMENTATION 61<br />

<strong>The</strong> subgrid values are updated within the main iteration loop of the CFD<br />

code. After 〈U〉, k, and ε or ω in the main grid have been updated in one<br />

iteration, the subgrid update function is called. This function is essentially<br />

the main iteration loop of another, embedded CFD code that performs one<br />

iteration <strong>to</strong> solve for 〈U〉, k, and ˜ε or ω in the subgrid. Before the subgrid<br />

calculations are performed, data are taken from the main grid <strong>to</strong> act as<br />

boundary conditions on the subgrid. After the subgrid is updated, subgrid-<br />

averaged values are extracted in order <strong>to</strong> act as wall function inputs <strong>to</strong> the<br />

next main grid iteration. Thus, there is a cyclic exchange of information.<br />

<strong>The</strong> data required as input <strong>to</strong> the subgrid calculation are the values of 〈U〉,<br />

k, and ε or ω at the outer extent of the subgrid (where the subgrid ends and<br />

the main grid transport equations begin). Transport equations are solved in<br />

every subgrid cell (except in the case of the k-ω model, where the very-near-<br />

wall values of ω are prescribed, as discussed above). At the wall node, zero<br />

values of 〈U〉, k, and ˜ε are set. At the node placed farthest from the wall,<br />

at the outer-most point on the subgrid, the values of 〈U〉, k, and ˜ε or ω are<br />

set as being equal <strong>to</strong> the corresponding values linearly interpolated from the<br />

two nearest main grid nodes.<br />

<strong>The</strong> boundary conditions on the main grid are more complex. 〈U〉 receives its<br />

boundary condition as in the above section on log-law wall functions, except<br />

that the wall shear stress, τw is obtained from the application of New<strong>to</strong>n’s<br />

law of viscosity at the subgrid’s near-wall cell node:<br />

<br />

∂ 〈U〉 <br />

τw = ρν <br />

∂y<br />

y=0<br />

Taking s1 as the near-wall subgrid cell node and discretising ∂〈U〉<br />

∂y<br />

<br />

<br />

y=0<br />

(4.28)<br />

within


CHAPTER 4. NUMERICAL IMPLEMENTATION 62<br />

the subgrid, SU of the main grid near-wall cell node is modified as follows:<br />

SU = S ′ <br />

〈U〉s1<br />

U − ν<br />

(4.29)<br />

To obtain the main grid boundary condition on k, volume-weighted subgrid<br />

averages of turbulent production (Pk) and dissipation (ε = ˜ε+ˆε) are required.<br />

<strong>The</strong>se are readily obtained after the subgrid governing equations have been<br />

solved. <strong>The</strong>se values may be integrated with respect <strong>to</strong> the main grid near-<br />

wall cell volume and then incorporated as a source in the main grid near-wall<br />

cell equation for k:<br />

For the k-ω model<br />

ys1<br />

SU = S ′ U + Pk − ε (4.30)<br />

ε = (ωk)β ∗<br />

(4.31)<br />

Note that, in the above equation for ε, the quotient is averaged. If the<br />

individual terms are averaged separately and then multiplied, the boundary<br />

condition will not be correct. Terms may only be averaged separately where<br />

the net source applied <strong>to</strong> an equation is a linear combination of those terms.<br />

<strong>The</strong> boundary condition on ε is obtained in a very similar manner <strong>to</strong> the<br />

boundary condition on k. Rather than prescribing a value of ε in the main<br />

grid near-wall cell, the terms leading <strong>to</strong> the production and distruction of ˜ε<br />

are averaged throughout the subgrid, leading <strong>to</strong> a net production of ˜ε <strong>to</strong> be<br />

applied <strong>to</strong> the main grid. (In the main grid node, ˜ε may be taken as being<br />

synonymous with ε, since ˆε → 0 away from solid boundaries.) <strong>The</strong> modified<br />

source term applied <strong>to</strong> the main grid near-wall cell is<br />

SU = S ′ U + Cε1<br />

<br />

f1<br />

<br />

˜ε<br />

Pk − Cε2 f2<br />

k<br />

<br />

˜ε 2<br />

− E − Y (4.32)<br />

k


CHAPTER 4. NUMERICAL IMPLEMENTATION 63<br />

Unfortunately, a similar approach was found <strong>to</strong> be impossible in specifying<br />

the boundary condition on ω. This is thought <strong>to</strong> be due <strong>to</strong> the near-wall be-<br />

haviour of ω. Because ω approaches infinity near the wall, subgrid-averaged<br />

production and destruction values are difficult <strong>to</strong> obtain. <strong>The</strong>refore, ω is<br />

specified at the main grid near-wall cell node according <strong>to</strong> linear interpola-<br />

tion between the two closest subgrid nodes. Given an interpolated subgrid<br />

ω value, ωsg, the source terms on ω in the main grid are<br />

where G is a large number.<br />

SP = −G<br />

SU = G · (ωsg) (4.33)<br />

<strong>The</strong> <strong>UMIST</strong>-N subgrid wall function simplifies near-wall calculation, as com-<br />

pared <strong>to</strong> a low-Reynolds-number treatment, by removing pressure correction<br />

and the wall-normal momentum equation from the solution procedure. Thus,<br />

the flow calculation is reduced <strong>to</strong> a parabolic problem. Since PASSABLE<br />

is a parabolic code with wall-normal velocity constrained <strong>to</strong> be zero, the<br />

differences between the subgrid wall function approach and a low-Reynolds-<br />

number treatment of steady channel flow may be expected <strong>to</strong> be less sig-<br />

nificant, both in terms of accuracy and computational cost, than would be<br />

found in other flow geometries.<br />

One key difference remaining between the subgrid and low-Reynolds-number<br />

calculations is the necessity of applying subgrid results <strong>to</strong> the main grid as<br />

wall-function-type boundary conditions, rather than solving the governing<br />

equations in an uninterrupted manner through the near-wall region. Fur-<br />

thermore, a potential exists for instability <strong>to</strong> be introduced by the lag that<br />

exists between a main grid update and a subgrid update. This lag comes


CHAPTER 4. NUMERICAL IMPLEMENTATION 64<br />

about by the way in which main grid and subgrid solutions are updated in<br />

alternation, with data from each being passed as boundary conditions <strong>to</strong> the<br />

other, in turn. Finally, the subgrid allows a very different level of grid re-<br />

finement <strong>to</strong> be employed within the subgrid than in the main grid. Thus,<br />

extensive near-wall grid refinement may be achieved without the numerical<br />

problems that are usually associated with a single grid employing a very large<br />

expansion fac<strong>to</strong>r.<br />

4.5 Under-Relaxation<br />

Under-relaxation is a technique that involves hindering the rate of conver-<br />

gence of the CFD code in order <strong>to</strong> dampen spurious effects and reduce the<br />

risk of overshooting the converged solution.<br />

Generally, after an updated solution <strong>to</strong> one of the governing differential equa-<br />

tions has been obtained, the resulting calculated local values of a quantity φ<br />

will replace the previous values, according <strong>to</strong><br />

φ new = φ calc<br />

(4.34)<br />

When under-relaxation is used, an under-relaxation fac<strong>to</strong>r, α may be applied,<br />

and the updated values of φ will be the weighted average of φ calc and φ old ,<br />

according <strong>to</strong><br />

φ new = αφ calc + (1 − α) φ old<br />

(4.35)


CHAPTER 4. NUMERICAL IMPLEMENTATION 65<br />

In the discretised transport equations, under-relaxation is implemented by<br />

(AP ) ur = 1<br />

α AP<br />

(SU) ur = SU + (1 − α) AP φp<br />

(4.36)<br />

<strong>The</strong> under-relaxation fac<strong>to</strong>rs used in this thesis are shown in Tables 4.2 and<br />

4.3.<br />

Table 4.2: Under-relaxation fac<strong>to</strong>rs used in the main grid<br />

〈U〉 0.8<br />

k 0.5<br />

ε 1.0<br />

˜ε 1.0<br />

ω 1.0<br />

Table 4.3: Under-relaxation fac<strong>to</strong>rs used in the subgrid<br />

〈U〉 0.7<br />

k 0.7<br />

˜ε 0.7<br />

ω 0.7


Chapter 5<br />

Results<br />

<strong>The</strong> performance of the <strong>UMIST</strong>-N near-wall subgrid treatment was tested<br />

in channel flow. <strong>The</strong> PASSABLE code was used <strong>to</strong> implement a parabolic<br />

solution method <strong>to</strong> solve this flow domain. For this work, the PASSABLE<br />

code was extended <strong>to</strong> allow high-Reynolds-number (wall function and sub-<br />

grid) solutions. Also, a separate implementation of the turbulence models<br />

and a separate solver were written <strong>to</strong> be employed in solving the subgrid<br />

region near the wall (extending from y/δ = 0 <strong>to</strong> y/δ = 0.2).<br />

<strong>The</strong> models implemented in the subgrid are the standard k-ε model [33] with<br />

Yap correction [92] and the 1988 k-ω model [89]. PASSABLE supports a<br />

range of turbulence models for the solution of the main grid. <strong>The</strong> models<br />

employed in this work were the low-Reynolds-number k-ε model with Yap<br />

correction, the 1988 k-ω model, and the high-Reynolds-number k-ε model<br />

(newly added). <strong>The</strong> configurations of turbulence models used in this work<br />

are summarised descriptively in Table 5.1.<br />

66


CHAPTER 5. RESULTS 67<br />

Table 5.1: Configurations of turbulence models<br />

Low-Re k-ε <strong>The</strong> subgrid was not employed and the low-Reynolds-<br />

number standard k-ε model was used.<br />

k-ε with log law <strong>The</strong> subgrid was not employed and the high-Reynolds-<br />

number standard k-ε model was used with a<br />

log-law boundary condition near the wall.<br />

Subgrid k-ε <strong>The</strong> high-Reynolds-number standard k-ε model<br />

was used in the main grid with a near-wall boundary<br />

condition derived from a subgrid solution. <strong>The</strong> subgrid<br />

employed the low-Reynolds-number standard k-ε model.<br />

Subgrid k-ω <strong>The</strong> 1988 k-ω model was used in the main grid with a<br />

near-wall boundary condition derived from a subgrid<br />

solution. <strong>The</strong> subgrid also employed the 1988 k-ω model.<br />

Three test cases were implemented: steady channel flow, channel flow driven<br />

by a periodically variable pressure gradient, and channel flow constrained <strong>to</strong><br />

exhibit a periodically variable bulk flow rate. <strong>The</strong> two periodic cases differ<br />

in that the former matches the pressure gradient fluctuation of the DNS<br />

study of Kawamura & Homma [29] while the latter matches the bulk flow<br />

variation presented in that DNS result. <strong>The</strong> steady flow case was performed<br />

at a nominal friction velocity Reynolds number of Reτ = 180. 1 <strong>The</strong> periodic<br />

flow cases were configured such that the time average flow would conform <strong>to</strong><br />

1 <strong>The</strong> boundary conditions on the code were configured such that 180 would be a correct<br />

value of Reτ based on a theoretical analysis. <strong>The</strong> actual Reynolds number obtained was<br />

different, as discussed below. This is what is meant by ‘nominal’ Reτ .


CHAPTER 5. RESULTS 68<br />

a nominal Reτ of 180.<br />

5.1 Steady <strong>Flow</strong> Results<br />

Although the principal goal of this work is <strong>to</strong> test the performance of the<br />

<strong>UMIST</strong>-N method in periodic flow, the steady flow case offers a simpler test<br />

case in which <strong>to</strong> verify the code.<br />

Steady channel flow results for the k-ε and k-ω models are widely available.<br />

<strong>The</strong> <strong>UMIST</strong>-N approach with the standard k-ε model was applied <strong>to</strong> channel<br />

flow using the TEAM code by Gant during the early development of <strong>UMIST</strong>-<br />

N [21]. <strong>The</strong> application of the <strong>UMIST</strong>-N approach with a k-ω model <strong>to</strong><br />

steady channel flow is new.<br />

All the model configurations shown in Table 5.1 were applied <strong>to</strong> steady chan-<br />

nel flow and the results are presented here. Figures 5.1 & 5.2 show U + and<br />

k + (respectively) plotted against y/δ. Figures 5.3 & 5.4 show a better view<br />

of near-wall values plotted against y + on logarithmic scales. <strong>The</strong> results are<br />

compared the analytical profiles from Chapter 3 and <strong>to</strong> the data of Kim et<br />

al. [30], who employed the same Reτ.<br />

<strong>The</strong> pressure gradient driving the flow was chosen <strong>to</strong> produce a nominal Reτ<br />

of 180. By applying the definition of Reτ (Equation 3.25), a value of Uτ was<br />

obtained that was used in nondimensionalisation. From Equations 3.23 &<br />

3.19, a corresponding<br />

∂〈P 〉<br />

∂x<br />

was specified. This ∂〈P 〉<br />

∂x<br />

was allowed <strong>to</strong> drive the<br />

flow, with the actual calculated flow rate being a function of the turbulence<br />

model used.


CHAPTER 5. RESULTS 69<br />

<strong>The</strong> solution procedure employed involved initialising flow variables <strong>to</strong> zero,<br />

applying the driving pressure gradient, and iteratively updating the solution<br />

until convergence was achieved. Gradients in time ( ∂<br />

∂t<br />

terms) were included<br />

in the solution, so that unconverged results may have been regarded con-<br />

ceptually as the transient response of the flow <strong>to</strong> a step change in pressure.<br />

However, transient results are not shown. ∂<br />

∂t<br />

terms do not affect the con-<br />

verged result in steady flow and were included primarily for debugging.<br />

In Figure 5.1, the k-ε solutions overpredict peak velocity and bulk flow, while<br />

the k-ω model underpredicts. Figure 5.3 shows that the 〈U〉 + results nearer<br />

<strong>to</strong> the wall are similar for each model and match the DNS result closely.<br />

Equation 3.17 can be evaluated at y = 0 <strong>to</strong> obtain<br />

<br />

d 〈U〉 y=0<br />

τw = ρν<br />

dy<br />

d〈P 〉<br />

dx by<br />

Recalling Equation 3.19, τw is related <strong>to</strong><br />

<br />

d 〈P 〉<br />

τw = −δ<br />

dx<br />

(5.1)<br />

(5.2)<br />

<strong>The</strong>refore, close agreement on 〈U〉 + near the wall is <strong>to</strong> be expected from the<br />

prescription of<br />

d〈P 〉<br />

dx . <strong>The</strong> deviation in 〈U〉+ occurs farther from the wall and<br />

is likely due <strong>to</strong> the prediction of νt.<br />

<strong>The</strong> subgrid solution produces results that are very similar <strong>to</strong> those of the<br />

low-Reynolds-number approach. This is <strong>to</strong> be expected, since the internal<br />

subgrid calculation is identical <strong>to</strong> a standard low-Reynolds-number calcu-<br />

lation in a parabolic solution scheme. <strong>The</strong> unique aspect of the subgrid<br />

approach as compared <strong>to</strong> a low-Reynolds-number parabolic solution is in the<br />

transferral of information between the subgrid and the main grid via mutu-<br />

ally dependent boundary conditions. However, this complexity is unlikely <strong>to</strong>


CHAPTER 5. RESULTS 70<br />

affect the solution of steady channel flow. This assumption is supported by<br />

the smoothness of the subgrid solution on both U + and k + in the vicinity<br />

of y/δ = 0.2, where the subgrid and the main grid meet. <strong>The</strong> small differ-<br />

ence observed between the subgrid and low-Reynolds-number k-ε solutions<br />

is likely due <strong>to</strong> differences in grid refinement. Although the subgrid solution<br />

uses fewer cells in <strong>to</strong>tal (summing the subgrid and the main grid cells) than<br />

the low-Reynolds-number approach, it offers enhanced near-wall refinement<br />

within the subgrid, with a coarser grid employed away from the wall. <strong>The</strong><br />

fact that the subgrid solution produced a velocity profile that is very slightly<br />

nearer <strong>to</strong> the DNS result highlights a secondary benefit of the <strong>UMIST</strong>-N<br />

approach, which is that the application of boundary conditions at a subgrid<br />

/ main grid boundary allows an opportunity for extensive and sudden vari-<br />

ation in grid refinement without some of the numerical challenges usually<br />

associated with inconsistent cell sizes.<br />

<strong>The</strong> logarithmic law of the wall produces excellent results in steady channel<br />

flow. This is <strong>to</strong> be expected, since it was designed and tuned with this par-<br />

ticular flow configuration in mind. Figures 5.1 & 5.3 indicate nearly perfect<br />

agreement on velocity. <strong>The</strong> discrepancy in k + seen in Figures 5.2 & 5.4 is<br />

likely the result of that variable’s definition within the k-ε model. Although<br />

k is intended <strong>to</strong> model turbulent kinetic energy, the model constants have<br />

been chosen with the primary aim of producing good agreement with exper-<br />

iments in the prediction of velocities. k is related <strong>to</strong> velocity via the EVM.<br />

Functionally, then, it may be convenient <strong>to</strong> conceptualise k not as turbulent<br />

kinetic energy but simply as that parameter which is required by the EVM<br />

<strong>to</strong> produce desirable values of velocity. With this in mind, the log law’s pre-<br />

diction of k + near the wall in channel flow must match a hypothetical ideal<br />

k-ε solution that reproduces the DNS result for U + . It must not necessarily


CHAPTER 5. RESULTS 71<br />

match the DNS result for k + .<br />

<strong>The</strong> subgrid approach with the k-ω turbulence model performs rather well. It<br />

can be seen <strong>to</strong> underpredict U + in Figures 5.1 & 5.3, although by a relatively<br />

small margin. Away from the wall, the value of k in the k-ω model offers a<br />

good approximation <strong>to</strong> the DNS result, as seen in Figures 5.2 & 5.4.<br />

All of the solutions underpredict turbulent kinetic energy near the wall, as<br />

is particularly apparent in Figure 5.4. Tracing from the symmetry plane of<br />

the channel <strong>to</strong>wards the wall, the point at which the modelled solutions tend<br />

<strong>to</strong> depart from the DNS result is y + ≈ 40. This corresponds roughly <strong>to</strong> the<br />

outer extent of the buffer layer (see Table 2.4).<br />

Overall, the steady channel flow results appear consistent with prior work<br />

and offer credibility <strong>to</strong> the following calculations involving periodic flow.<br />

5.2 Prescribed <strong>Periodic</strong> Pressure Gradient<br />

<strong>The</strong> flow was subjected <strong>to</strong> the same periodically variable pressure gradient<br />

employed by Kawamura & Homma [29] in their DNS study. This pressure<br />

gradient was defined by<br />

∂ +<br />

<br />

〈P 〉<br />

2π<br />

= 1 + 6 sin<br />

∂x<br />

6 t+<br />

<br />

where<br />

∂〈P 〉<br />

∂x<br />

(5.3)<br />

and t (time) are nondimensionalised by<br />

+<br />

∂ 〈P 〉<br />

∂x<br />

=<br />

<br />

∂〈P 〉<br />

∂x<br />

2<br />

ρ(Uτ ) ss<br />

(5.4)<br />

t + = (Uτ) ss t<br />

δ<br />

δ<br />

(5.5)


CHAPTER 5. RESULTS 72<br />

(Uτ) ss is calculated from (Reτ) ss = 180. This implies that<br />

+<br />

∂〈P 〉<br />

∂x<br />

ss<br />

= 1.<br />

Because of Equation 5.2, this also suggests a time-mean value of τw corre-<br />

sponding <strong>to</strong> the steady flow case considered above.<br />

To generate the periodic flow results, a converged steady flow solution was<br />

first reached. <strong>The</strong>n, pressure gradient variation was introduced according <strong>to</strong><br />

Equation 5.3. To ensure convergence, 1000 time steps were used per period in<br />

the periodic flow case. At each time step, the code was run until convergence<br />

was achieved. However, this convergence was intermediate in the sense that<br />

it represented an estimate of the solution using the given ∂<br />

∂t<br />

information. To<br />

obtain results representing the effect of long-term periodicity, this process was<br />

repeated through several periods until only negligible changes were observed<br />

in subsequent periods. Thus, start-up effects were eliminated.<br />

Figures 5.5 & 5.6 show the prescribed pressure gradient, the calculated bulk<br />

flow rate, and the calculated wall shear stress plotted with DNS results. <strong>The</strong><br />

graphs show values plotted from periods 0 <strong>to</strong> 2. This is not meant <strong>to</strong> imply<br />

that the first two periods of generated output are plotted. <strong>The</strong>se graphs are<br />

produced from a final, converged period plotted twice for visual clarity.<br />

Figure 5.5 shows a marked underprediction of the amplitude of variation<br />

in bulk flow for all of the models employed. Figure 5.6 also indicates that<br />

wall shear stress is underpredicted. This is consistent with the findings of<br />

Addad [1], and, as he pointed out, the relationship between τw, U, and<br />

is manifest in the x-momentum equation governing the flow (Equation 3.5).<br />

Employing the definition of the shear stress, τ(y) from Equation 3.17, the<br />

x-momentum equation is<br />

∂ 〈U〉<br />

∂t<br />

∂ 〈P 〉<br />

= −1<br />

ρ ∂x<br />

1 dτ<br />

+<br />

ρ dy<br />

∂〈P 〉<br />

∂x<br />

(5.6)


CHAPTER 5. RESULTS 73<br />

Integrating with respect <strong>to</strong> y,<br />

δ <br />

∂<br />

1 ∂ 〈P 〉<br />

〈U〉 dy = −<br />

y|<br />

∂t<br />

ρ ∂x<br />

y=δ<br />

<br />

τ <br />

y=0 + <br />

ρ<br />

0<br />

τ=0<br />

τ=τw<br />

Dividing by δ and noting that 1<br />

δ<br />

〈U〉 dy = U, Equation 5.7 becomes<br />

δ 0<br />

dU<br />

dt<br />

+ τw<br />

ρδ<br />

∂ 〈P 〉<br />

= −1<br />

ρ ∂x<br />

(5.7)<br />

(5.8)<br />

Thus, in unsteady channel flow, the driving pressure gradient applies energy<br />

<strong>to</strong> both overcome a wall shear stress and <strong>to</strong> accelerate the flow. <strong>The</strong> fact that<br />

peaks and troughs in U are underpredicted in Figure 5.5 indicates that dU<br />

dt is<br />

underestimated by the turbulence models. τw also displays a lower amplitude<br />

of variation than expected, as seen in Figure 5.6. This is consistent with<br />

<br />

<br />

(Equation 5.1). Together, these<br />

Figure 5.5, since τw is a function of ∂〈U〉<br />

∂y<br />

results suggest that the energy is added <strong>to</strong> and removed from the flow at a<br />

less than realistic rate.<br />

<strong>The</strong> equality in Equation 5.8 appears <strong>to</strong> be compromised when using these<br />

turbulence models in periodic flow. In terms of the x-momentum equation<br />

(Equation 3.5), this suggests that the models are not as well tuned <strong>to</strong> predict<br />

〈uv〉 in periodic channel flow as in steady channel flow. Errors in predicting<br />

〈uv〉 produce deviations in 〈U〉, particularly farther from the wall, as seen in<br />

Figures 5.11 & 5.12. This effects the dU<br />

dt<br />

y=0<br />

term in Equation 5.8.<br />

Figures 5.7, 5.8, 5.9 & 5.10 show flow variables plotted as a function of phase<br />

angle at various locations throughout the channel (y/δ = 0.1, 0.2, 0.5 & 0.9,<br />

respectively).<br />

Results for the k-ε log law do not appear in Figure 5.7, because y/δ = 0.1<br />

is outside the calculated flow field when the log law is used. In Figure 5.8,


CHAPTER 5. RESULTS 74<br />

where y/δ = 0.2, the log law produces results for k + that are in phase with the<br />

results for 〈U〉 + . This is <strong>to</strong> be expected from the log law equations. Further<br />

from the wall, where the k-ε model equations are applied, the generated<br />

result is improved, but still offers little encouragement as <strong>to</strong> the applicability<br />

of the log law <strong>to</strong> this flow.<br />

Examining turbulent kinetic energy, k + in Figures 5.7, 5.8, 5.9 & 5.10, it<br />

appears that the subgrid k-ε treatment closely matches the low-Reynolds-<br />

number k-ε solution for y/δ ≤ 0.2, within the subgrid region. Outside of the<br />

subgrid region (Figures 5.9 & 5.10), the subgrid solution appears <strong>to</strong> capture<br />

the effects of variations in<br />

∂〈P 〉<br />

∂x<br />

slightly less effectively than the low-Reynolds-<br />

number counterpart. This suggests that the process of averaging the subgrid<br />

solution <strong>to</strong> be applied as a boundary condition <strong>to</strong> the main grid introduces<br />

some discernible measure of inaccuracy. <strong>The</strong> subgrid solution can therefore<br />

produce different results than a calculation that is continuous throughout the<br />

flow field (standard low-Reynolds-number treatment), even when the same<br />

turbulence model is used in each case.<br />

<strong>The</strong> k-ω solution produces and amplitude of oscillation of τw that more closely<br />

matches the DNS than that produced by the k-ε subgrid solution (Figure<br />

5.6). Correspondingly, the k-ω model produces fluctuations in bulk velocity<br />

that are also slightly improved (Figure 5.5). This is an indication of the<br />

applicability of the k-ω model <strong>to</strong> boundary layer flow. <strong>The</strong> k-ω model exhibits<br />

a less prominent peak in turbulent kinetic energy than is produced by the<br />

k-ε model (Figures 5.7, 5.8, 5.9 & 5.10). This may be compounded by the<br />

general tendency of the k-ω model <strong>to</strong> offer lower values of k, as observed in<br />

the steady flow results.


CHAPTER 5. RESULTS 75<br />

Figures 5.11, 5.12, 5.13 & 5.14 offer snapshots showing 〈U〉 + and k + , respec-<br />

tively, throughout the channel at various discrete phase positions. Figures<br />

5.15, 5.16, 5.17 & 5.18 show the same again, but plotted against y + rather<br />

than y/δ.<br />

Because of the general inaccuracy of all the models, it is difficult <strong>to</strong> make<br />

comparisons based on Figures 5.11 <strong>to</strong> 5.18. Figures 5.13 & 5.14 again high-<br />

light the tendency of the log law solution of k + <strong>to</strong> generate spurious results<br />

at y/δ = 0.2 where the log law is applied. Furthermore, Figures 5.17 & 5.18<br />

show that all the models produced an underprediction of k + near the wall.<br />

This is similar <strong>to</strong> the underprediction observed in the steady flow case.<br />

To better facilitate comparisons between the models, and following the pe-<br />

riodic channel flow investigation of Addad [1], a further test case was con-<br />

sidered in which the modelled flows are constrained <strong>to</strong> exhibit the same bulk<br />

flow rates shown in the DNS study. <strong>The</strong>se results are presented below.<br />

5.3 Prescribed <strong>Periodic</strong> Bulk <strong>Flow</strong> Rate<br />

<strong>The</strong> code was configured <strong>to</strong> match the bulk flow from the DNS study of<br />

Kawamura & Homma [29], rather than the<br />

∂〈P 〉<br />

∂x<br />

data. Figure 5.19 shows<br />

the resulting pressure variations compared <strong>to</strong> the pressure variation in the<br />

DNS study. As expected, the resultant pressure variation is higher. <strong>The</strong><br />

amplitude on the fluctuation in<br />

∂〈P 〉<br />

∂x is greater than the DNS amplitude by<br />

approximately 100%. Correspondingly, wall shear stress is increased.<br />

As seen in Figure 5.20, the k-ε low-Reynolds-number and subgrid results now


CHAPTER 5. RESULTS 76<br />

more closely resemble the DNS data for τw, except that they exhibit a double-<br />

peak behaviour. τw produced by the k-ω subgrid matches the troughs in the<br />

DNS data effectively, but the peak values are disappointingly overestimated.<br />

<strong>The</strong> asymmetry in τw suggests that the k-ω model predicts excessively steep<br />

profiles of 〈U〉 + near the wall for most phase angles, as can be seen in Figures<br />

5.31 & 5.32.<br />

<strong>The</strong> further figures mirror the presentation format employed in the previous<br />

section. Figures 5.21, 5.22, 5.23 & 5.24 display flow variables through time<br />

at y/δ locations of 0.1, 0.2, 0.5 & 0.9, respectively. Figures 5.25, 5.26, 5.27<br />

& 5.28 show snapshots of the channel at various phase angles, and Figures<br />

5.29, 5.30, 5.31 & 5.32 show these against a logarithmic y + scale.<br />

Figures 5.21, 5.22, 5.23 & 5.24 show a tendency for the low-Reynolds-number<br />

k-ε model <strong>to</strong> exhibit a delay in predicting an increase in k + when the flow<br />

is accelerating. When it appears (for example, at a phase angle of 7π/6<br />

in Figure 5.23), the anticipated increase in k + is sudden and sharp. At<br />

y/δ ≤ 0.5, the peak value of k + through time predicted by the low-Reynolds-<br />

number k-ε model (and indeed by all of the models considered in this work) is<br />

lower than the DNS result, but k + is actually overpredicted throughout most<br />

of the portion of the cycle during which the flow is decelerating. In Figure<br />

5.24, with y/δ = 0.9, the peak value of k + is overpredicted by all except the<br />

k-ω subgrid model and the overprediction of k + during deceleration of the<br />

flow is very pronounced in all of the modelled results.<br />

<strong>The</strong> use of the <strong>UMIST</strong>-N subgrid approach with the k-ε model does not ap-<br />

pear <strong>to</strong> alter the phase position of the suden increase in k + noted above. How-<br />

ever, the subgrid approach does appear <strong>to</strong> soften this effect. <strong>The</strong> maximum


CHAPTER 5. RESULTS 77<br />

gradient of k + seen in Figures 5.23 & 5.24 is lower than for the low-Reynolds-<br />

number k-ε model. Furthermore, the extent of over- and underprediction of<br />

k + is diminished throughout the cycle. One reason for this may be that<br />

diffusion is not accounted for across the subgrid boundary. Although the<br />

subgrid is constrained <strong>to</strong> match the main grid value of k + at y/δ = 0.2, there<br />

is no diffusion of k in<strong>to</strong> the main grid from the subgrid. <strong>The</strong> only information<br />

transmitted from the subgrid <strong>to</strong> the main grid in the calculation of the model<br />

equation of k is subgrid-averaged Pk and ε.<br />

Two other potential sources of discrepancy between subgrid and low-Reynolds-<br />

number treatments are detalied here. Firstly, greater numerical errors may<br />

occur in the subgrid solution as a result of the averaging process that is used<br />

<strong>to</strong> apply subgrid results <strong>to</strong> the main grid as boundary conditions. Secondly,<br />

the subgrid solution may be configured <strong>to</strong> offer greater near-wall grid refine-<br />

ment because of the fact that the subgrid cell sizes may be set independently<br />

of the main grid cell sizes. In this work, the subgrid solution did employ<br />

smaller cells near the wall than were used in the low-Reynolds-number treat-<br />

ment, because the use of a high-Reynolds-number model in the main grid<br />

allowed the use of fewer nodes in <strong>to</strong>tal. However, suitable grid refinement<br />

was employed in all cases so that the computed results may be assumed <strong>to</strong><br />

be independent of cell size. It must be highlighted that the potential of<br />

the subgrid treatment <strong>to</strong> achieve greater near-wall grid refinement without<br />

impacting the refinement of the main grid is a powerful feature of <strong>UMIST</strong>-N.<br />

Notably, any difference between the subgrid and low-Reynolds-number k-ε<br />

profiles is less discernable in Figures 5.21 & 5.22, where y/δ ≤ 0.2. This<br />

region falls within the subgrid itself, and the calculation there is substan-<br />

tially similar <strong>to</strong> a standard low-Reynolds-number treatment. <strong>The</strong> difference


CHAPTER 5. RESULTS 78<br />

between the subgrid and standard low-Reynolds-number treatments is in the<br />

averaging used within the subgrid <strong>to</strong> apply subgrid results as a boundary con-<br />

dition <strong>to</strong> the main grid. <strong>The</strong>refore, the impact of using a subgrid treatment<br />

manifests itself beyond y/δ = 0.2.<br />

<strong>The</strong> subgrid k-ω model predicts k + admirably well in Figures 5.21, 5.22, 5.23<br />

& 5.24. It does not exhibit the sudden increase in k + noted above, and actual<br />

predicted values of k + appear <strong>to</strong> be at least as accurate as those of other<br />

models at each of the traverse points shown. However, careful examination<br />

of the DNS data suggests that a slightly greater slope of k + with time is<br />

<strong>to</strong> be expected when k + is increasing than when it is decreasing. At lower<br />

values of y/δ, the k-ω model appears <strong>to</strong> fail <strong>to</strong> capture this effect, suggesting<br />

a qualitative error. Such a criticism would be harsh, however, in light of<br />

Figure 5.24 (y/δ = 0.9), in which the k-ω model does exhibit a greater slope<br />

on the increase of k + than on the decrease. On the whole, the model appears<br />

<strong>to</strong> offer good qualitative and quantitative accuracy.<br />

<strong>The</strong> favourable performance of the k-ω subgrid in predicting k + is most ap-<br />

parent at large y + , as seen in Figures 5.27, 5.28, 5.31 & 5.32. <strong>The</strong> k-ω<br />

prediction matches the DNS almost perfectly for y = δ and phase angles of<br />

5π/4 and 3π/2, and the model produces good results at y = δ for other phase<br />

angles. <strong>Near</strong> the wall, the behaviour of k + is more accurately reproduced by<br />

the k-ω subgrid for phase angles of π/2 and 3π/4 (favourable pressure gradi-<br />

ent) than by the k-ε subgrid. However, the k-ω model’s near-wall predictions<br />

are less favourable at phase angles of 3π/2 and 7π/4, under the influence of<br />

an adverse pressure gradient. At phase angles of 0 and π, there is little<br />

discernible difference among the near-wall predictions of the various models.


CHAPTER 5. RESULTS 79<br />

<strong>The</strong> log law offers predictions of k + in Figures 5.23 & 5.24 that are reasonable.<br />

<strong>The</strong> log law solution has also been improved greatly by the prescription of U<br />

rather than<br />

∂〈P 〉<br />

. It must once again be noted that the log law only produces<br />

∂x<br />

interesting results away from y/δ = 0.2, where it is applied for this particular<br />

flow.<br />

A ‘kinked’ region of large change in gradient tends <strong>to</strong> exist in many of the<br />

profiles of 〈U〉 + . This phenomenon can be seen most clearly in Figures 5.25<br />

& 5.26. It is produced by all of the models except for the log law treatment,<br />

and is particularly prevalent in the profiles produced by the low-Reynolds-<br />

number k-ε model. A kink indicates an underprediction of the diffusion of<br />

〈U〉 + within the vicinity of the kink. Deferring <strong>to</strong> the assumptions of the<br />

EVM, this would suggest a local underprediction of k + . This corollary can<br />

in fact be seen in the figures. Comparing phase angles of 3π/4, π, and 5π/4,<br />

suggests that the location of a kink in 〈U〉 + corresponds <strong>to</strong> a location of<br />

maximum underprediction of k + . This may be seen by examining Figures<br />

5.25 & 5.26 for velocity and 5.27 & 5.28 for k + .<br />

<strong>The</strong> same phenomenon of a kinked velocity profile can be observed indirectly<br />

in Figures 5.21, 5.22, 5.23 & 5.24, where 〈U〉 + predicted by the low-Reynolds-<br />

number k-ε model appears <strong>to</strong> lead the DNS result in phase closer <strong>to</strong> the wall,<br />

whereas it lags the DNS in phase at higher y/δ. This apparent peculiarity is<br />

in fact a symp<strong>to</strong>m of the changing slope of the 〈U〉 + profile produced by the<br />

k-ε model at different phase angles.<br />

All of the models produce flatter velocity profiles than the DNS, as seen in<br />

Figures 5.25, 5.26, 5.29 & 5.30. This is particularly true just after a phase<br />

angle of π, when the flow is subjected <strong>to</strong> an adverse pressure gradient and a


CHAPTER 5. RESULTS 80<br />

sudden increase in k + occurs for y/δ ≥ 0.5 (Figures 5.23 & 5.24). When the<br />

pressure gradient diminishes and becomes adverse, the effect on velocity is<br />

expected <strong>to</strong> manifest itself first near the wall and then propagate outwards.<br />

This happens because the laminar layer near the wall contains less kinetic<br />

energy and therefore less momentum. All of the modelled results exhibit<br />

this preferential slowing of the fluid near the wall when the pressure gradient<br />

becomes adverse, but they do so later in phase and <strong>to</strong> a lesser extent than<br />

what is seen in the DNS data. This can be seen in Figures 5.25 & 5.26.<br />

It is interesting <strong>to</strong> compare the snapshots for the phase angles of π/2, 3π/4,<br />

and π found in Figures 5.29 & 5.30. <strong>The</strong> excessive flatness of the 〈U〉 +<br />

profiles is apparent in all of these snapshots, but particularly at the phase<br />

angle of π, where the flattening occurs at a lower value of y + for all the models<br />

than at the phase angle of π/2. Between these two, the snapshot at phase<br />

angle 3π/4 shows the subgrid k-ω model already flattening at lower y + , while<br />

the low-Reynolds-number k-ε model persists in flattening only at higher y + .<br />

<strong>The</strong> subgrid k-ε modelled results follow the low-Reynolds-number solution<br />

for lower y + , but the 〈U〉 + profile flattens at y/δ ≈ 0.2 earlier <strong>to</strong> join the<br />

subgrid k-ω results for higher y + . <strong>The</strong> point at which the subgrid solution<br />

flattens corresponds roughly <strong>to</strong> the outer extent of the subgrid. Thus, at this<br />

transitional snapshot in time, when the 〈U〉 + profile is undergoing a change<br />

of shape, the solutions appears <strong>to</strong> be particularly sensitive <strong>to</strong> the averaging<br />

procedure used <strong>to</strong> calculate main grid boundary condition from the subgrid<br />

solution.<br />

An interesting feature of the <strong>UMIST</strong>-N subgrid approach is highlighted by<br />

the k-ω prediction of k + at phase angles of π/2 and 3π/4. Here, the k +<br />

profiles predicted by the subgrid k-ω model exhibit a double peak (two local


CHAPTER 5. RESULTS 81<br />

maxima). This is particularly visible in Figure 5.31, but Figure 5.27 shows<br />

that the local minimum between these two peaks lies at y/δ = 0.2, the limit<br />

of the subgrid region and also the point where boundary conditions from the<br />

subgrid are applied <strong>to</strong> the main grid and visa versa. Close inspection of all<br />

the k + profiles produced by the k-ω subgrid reveals minor discontinuities at<br />

this y/δ location for other phase angles also. This is a result of the way<br />

in which the subgrid solution of ω was applied as a boundary condition <strong>to</strong><br />

the main grid. <strong>The</strong> <strong>UMIST</strong>-N approach <strong>to</strong> k and ε boundary conditions<br />

is <strong>to</strong> calculate subgrid averaged production and dissipation terms for these<br />

parameters, which are then applied as source terms within the near-wall cell<br />

in the main grid solution. However, ω tends <strong>to</strong> infinity at the wall, and<br />

source terms on ω become numerically unwieldy in near-wall cells. Thus,<br />

subgrid averaged production of ω could not be obtained and instead the<br />

subgrid profile of ω was interpolated <strong>to</strong> the main grid node and applied there<br />

as a fixed value. 2 This boundary condition appears <strong>to</strong> be less effective than<br />

the average source approach, and, until a more satisfac<strong>to</strong>ry solution <strong>to</strong> the ω<br />

boundary condition problem can be found, it advises against the applicability<br />

of the <strong>UMIST</strong>-N approach <strong>to</strong> a k-ω model solution. This aside, the subgrid<br />

k-ω results appear satisfac<strong>to</strong>ry when considered as a whole.<br />

In addition <strong>to</strong> the above difficulty in the prescription of an ω boundary<br />

condition, further difficulties were encountered in attempting <strong>to</strong> run the k-ω<br />

model in the subgrid with a high-Reynolds-number k-ε model in the main<br />

grid. Poor results were seen <strong>to</strong> arise from the difference in the values of k<br />

produced by the two models under similar conditions. In general, the k-<br />

ω model produces lower values of k than the k-ε model, as noted in the<br />

2 See Chapter 4 for more information.


CHAPTER 5. RESULTS 82<br />

steady flow results. In order <strong>to</strong> reconcile the two models, it was proposed<br />

that a blending function similar <strong>to</strong> that used by the SST model [42] may be<br />

required. <strong>The</strong> development effort was s<strong>to</strong>pped because the added complexity<br />

would have compromised the relevance of the <strong>UMIST</strong>-N approach.<br />

<strong>The</strong> <strong>UMIST</strong>-N approach has been successfully applied <strong>to</strong> periodic flow. <strong>The</strong><br />

subgrid calculation offers discernibly different results than the standard low-<br />

Reynolds-number k-ε model, but neither may readily be identified as supe-<br />

rior. <strong>The</strong>refore, the <strong>UMIST</strong>-N approach is applicable <strong>to</strong> periodic flow. <strong>The</strong><br />

potential breadth of applicability of the approach is further indicated by its<br />

application <strong>to</strong> a k-ω model solution, although the challenge of obtaining main<br />

grid boundary conditions with the k-ω model is highlighted by these results.


Chapter 6<br />

Conclusions & Suggestions for<br />

Future Work<br />

In the present study, the <strong>UMIST</strong>-N near-wall subgrid treatment has been<br />

applied <strong>to</strong> time-variant flow. <strong>The</strong> cases considered have been periodic and<br />

steady channel flow. Results have been matched <strong>to</strong> DNS data on driving<br />

pressure gradient and on bulk flow rate. <strong>The</strong> logarithmic law of the wall, the<br />

k-ε subgrid solution and the low-Reynolds-number k-ε solution were sub-<br />

stantially similar in the steady case, as expected. In the unsteady case, the<br />

subgrid produced results that were much improved over the log law, bear-<br />

ing more similarity <strong>to</strong> the low-Reynolds-number results. <strong>The</strong> subgrid and<br />

low-Reynolds-number results were discernibly different, however, and some<br />

of the features <strong>to</strong> be considered when implementing a subgrid treatment were<br />

highlighted.<br />

<strong>The</strong> use of the k-ω turbulence model within the <strong>UMIST</strong>-N subgrid was also<br />

83


CHAPTER 6. CONCLUSIONS & SUGGESTIONS FOR FUTURE WORK84<br />

tried in each of the above test cases. <strong>The</strong> model produced good results overall,<br />

but its performance was hindered by spurious behaviour at the interface<br />

between the subgrid and the main grid. Numerical challenges associated<br />

with the implementation of the k-ω model within a subgrid approach have<br />

been revealed during this work and were highlighted in the results.<br />

Further steady channel flow data from various sources were also compiled<br />

within this thesis and emperical profiles were presented <strong>to</strong> identify trends.<br />

<strong>The</strong>se profiles could serve as a <strong>to</strong>ol <strong>to</strong> CFD researchers who are engaged in<br />

early development or debugging.<br />

<strong>The</strong> <strong>UMIST</strong>-N approach has been shown <strong>to</strong> be applicable <strong>to</strong> periodic flow.<br />

<strong>The</strong> potential breadth of applicability of the approach is further indicated by<br />

its application <strong>to</strong> a k-ω model solution, despite challenges that were encoun-<br />

tered. Further investigation in<strong>to</strong> the usefulness of the <strong>UMIST</strong>-N approach<br />

in the computation of time-variant industrial flows appears justified.<br />

With particular reference <strong>to</strong> industrial applicability, one potential avenue<br />

for the further development of <strong>UMIST</strong>-N is <strong>to</strong>wards its use in Large Eddy<br />

Simulation. Most LES calculations rely on wall functions, and hybrid com-<br />

binations of RANS and LES are not unknown. <strong>The</strong>refore, a potential exists<br />

for <strong>UMIST</strong>-N <strong>to</strong> offer a benefit <strong>to</strong> LES. <strong>The</strong> satisfac<strong>to</strong>ry performance of the<br />

subgrid method in periodic flow within this study suggests that <strong>UMIST</strong>-N<br />

may be suitable when exposed <strong>to</strong> the rapid local oscillations that can be<br />

associated with the movement of large-scale turbulent structures in LES.<br />

As a <strong>to</strong>ol for industrial modelling, <strong>UMIST</strong>-N offers an intermediate choice,<br />

in terms of accuracy and computational cost, between a standard wall func-<br />

tion boundary and a low-Reynolds-number treatment. This choice could be


CHAPTER 6. CONCLUSIONS & SUGGESTIONS FOR FUTURE WORK85<br />

further optimised or simply expanded by the incorporation of other turbu-<br />

lence models within the subgrid. This work has applied the k-ω model <strong>to</strong> the<br />

subgrid. A further attempt was made <strong>to</strong> mix a k-ω subgrid with a k-ε main<br />

grid, but the development effort was s<strong>to</strong>pped when it became clear that dif-<br />

ferences in the predicted values of k would necessitate the use of a blending<br />

function similar <strong>to</strong> that used by the SST [42].<br />

Building upon the use of k-ω, future work could be directed <strong>to</strong>wards the<br />

use of simpler turbulence models, particularly a one-equation model and an<br />

algebraic model within the subgrid. If this can be done while retaining a<br />

k-ε treatment in the main grid, then the adapted subgrid would offer an<br />

additional level of flexibility in the cost/accuracy tradeoff associated with it.<br />

Furthermore, the <strong>UMIST</strong>-N wall function could be applied <strong>to</strong> a further va-<br />

riety of flows. Most notably, higher Reynolds numbers could be considered.<br />

<strong>The</strong>re is a difficulty in obtaining DNS results at higher Reynolds numbers,<br />

but perhaps the completion of this work will allow the subgrid approach <strong>to</strong><br />

be confidently compared <strong>to</strong> experiments for periodic flow at higher Reynolds<br />

numbers. Also, this work applies the subgrid treatment <strong>to</strong> periodic flow at<br />

a relatively low frequency of oscillation and without sufficient amplitude <strong>to</strong><br />

generate flow reversal. A higher frequency and a larger amplitude would<br />

provide an interesting test case, provided that suitable data can be found<br />

for comparison. Such a test case would advise as <strong>to</strong> the applicability of<br />

<strong>UMIST</strong>-N <strong>to</strong> the flow inside engines.<br />

In a two-dimensional grid, the <strong>UMIST</strong>-N method involves simultaneously<br />

s<strong>to</strong>ring subgrid results throughout the flow field. In addition <strong>to</strong> a s<strong>to</strong>rage<br />

cost, this presents an added complexity in the implementation of the method.


CHAPTER 6. CONCLUSIONS & SUGGESTIONS FOR FUTURE WORK86<br />

An attempt could be made <strong>to</strong> reduce this s<strong>to</strong>rage requirement and allow the<br />

subgrid treatment on a given main grid cell <strong>to</strong> run as a self-contained function.<br />

If this can be done without a prohibitive sacrifice of accuracy, it could further<br />

broaden the industrial usefulness of <strong>UMIST</strong>-N.


Bibliography<br />

[1] Addad, Y., Application of eddy viscosity and stress transport models<br />

in periodic channel flow, PhD thesis, University of Manchester Institute<br />

of Science and Technology (<strong>UMIST</strong>), 2000.<br />

[2] Ahmadi-Befrui, B., Arcoumanis, C., Bicen, A. F., Gossman, A. D., Ja-<br />

hanbakhsh, A., & Whitelaw, J. H., Calculations and measurements<br />

of the flow in a mo<strong>to</strong>red model engine and implications for open-<br />

chamber direct injection engines, ASME, Three-dimensional turbulent<br />

shear flows, pp. 1-9, 1982.<br />

[3] Amano, R. S., Development of a turbulence near-wall model and its ap-<br />

plication <strong>to</strong> separated and reattached flows, Numerical Heat Transfer,<br />

vol. 7, p. 59-75, 1984.<br />

[4] An<strong>to</strong>nia, R. A. & Kim, J., Low-Reynolds-number effects on near-wall<br />

turbulence, Journal of Fluid Mechanics, vol. 276, pp. 61-80, 1994.<br />

[5] Balaras, E., Benocci, C., & Piomelli, U., Two-layer approximate<br />

boundary condition large-eddy simulations, AIAA Journal, vol. 34, pp.<br />

1111-1119, 1996.<br />

87


BIBLIOGRAPHY 88<br />

[6] Boussinesq, J., Théorie de l’écoulement <strong>to</strong>urbillant, Mém. Prés. par<br />

div. savant à L’Academie Scientifique de Paris, vol. 23, p. 46, 1877.<br />

[7] Boyer, V. & Laurence, D., A shape function wall function approach<br />

for high and low Reynolds near-wall turbulence models, International<br />

Journal of Numerical Methods in Fluids, vol. 40, pp. 241-251, 2002.<br />

[8] Chapman, D. R. & Kuhn, G. D., <strong>The</strong> limiting behaviour of turbulence<br />

near a wall, Journal of Fluid Mechanics, vol. 170, pp. 265-292, 1986.<br />

[9] Chou, P. Y., On the velocity correlations and the solution of the equa-<br />

tions of turbulent fluctuation, Quarterly <strong>Applied</strong> Mathematics, vol. 3,<br />

p. 38, 1945.<br />

[10] Coakley, T. J., Turbulence modelling methods for the compressible<br />

Navier-S<strong>to</strong>kes equations, AIAA, no. 83-1693, 1983.<br />

[11] Comte-Bellot, G.,<br />

Écoulement turbulent entre deux parios paralleles,<br />

Publications Scientifiques et Techniques du Ministère de l’Air, vol. 419,<br />

1965.<br />

[12] Cot<strong>to</strong>n, M. A. & Ismael, J. O., An examination of periodic turbulent<br />

pipe flow using a low-Reynolds-number k-ε turbulence model, Pro-<br />

ceedings of the 8th Symposium on Turbulent Shear <strong>Flow</strong>s, Techhnical<br />

University of Munich, no. III-6, 1998.<br />

[13] Cot<strong>to</strong>n, M. A., Craft, T. J., Guy, A. W., & Launder, B. E., On mod-<br />

elling periodic motion with turbulence closures, <strong>Flow</strong>, Turbulence and<br />

Combustion, vol. 67, pp. 143-158, 2001.


BIBLIOGRAPHY 89<br />

[14] Craft, T. J., Launder, B. E., & Suga, K., Development and application<br />

of a cubic eddy-viscosity model of turbulence, International Journal of<br />

Heat and Fluid <strong>Flow</strong>, vol. 17, pp. 108-115, 1996.<br />

[15] Craft, T. J., Developments in a low-Reynolds-number second-moment<br />

closure and its application <strong>to</strong> separating and reattaching flows, Inter-<br />

national Journal of Heat and Fluid <strong>Flow</strong>, vol. 19, pp. 541-548, 1998.<br />

[16] Craft, T.J., Gant S.E., Iacovides, H., Launder B.E. & Robinson,<br />

C.M.E., Computational study of flow around the ”Ahmed” car body<br />

(case 9.4), 9th ERCOFTAC/IAHR/COST Workshop on Refined Tur-<br />

bulence Modelling, Darmstadt, Germany, 2001.<br />

[17] Craft, T.J., Gerasimov, A. V, Iacovides, H., and Launder B.E.,<br />

Progress in the generalization of wall-function treatments, Interna-<br />

tional Journal of Heat and Fluid <strong>Flow</strong>, vol. 23, pp. 148-160, 2002.<br />

[18] Craft, T.J., Gant S.E., Iacovides, H., and Launder B.E., A new wall<br />

function strategy for complex turbulent flows, Numerical Heat Transfer<br />

Part B: Fundamentals, vol. 45, no. 4, pp. 301-318, 2004.<br />

[19] Davidov, B. I., On the statistical dynamics of an incompressible tur-<br />

bulent fluid, Dokl. Akad. Nauk SSSR, vol. 136, pp. 47-50, 1987.<br />

[20] Ferziger, J. H., & Perić, M., Computational Methods for Fluid Dynam-<br />

ics, Springer, London, 2002.<br />

[21] Gant, S. E., Development and application of a new wall function for<br />

complex turbulent flows, PhD thesis, University of Manchester Institute<br />

of Science and Technology (<strong>UMIST</strong>), 2002.


BIBLIOGRAPHY 90<br />

[22] Hanjalic, K. & Launder, B. E., Contribution <strong>to</strong>wards a Reynolds-stress<br />

closure for low-Reynolds-number turbulence, Journal of Fluid Mechan-<br />

ics, vol. 74, pp. 593-610, 1976.<br />

[23] Harlow, F. H., & Nakayama, P. I., Transport of turbulence energy<br />

decay rate, Los Alamos Science Labora<strong>to</strong>ries, University of California<br />

Report, no. LA-3854, 1968.<br />

[24] Hino, M., Kashiwayanagi, M., Nakayama, A., & Hara, T., Experiments<br />

on the turbulence statistics and the structure of a reciprocating oscil-<br />

la<strong>to</strong>ry flow, Journal of Fluid Mechanics, vol. 131, pp. 363-400, 1983.<br />

[25] Iacovides, H. & Launder, B. E., PSL - an economical approach <strong>to</strong> the<br />

numerical analysis of near-wall elliptic flows, ASME Journal of Fluids<br />

Engineering, vol. 106, p. 241, 1984.<br />

[26] Jensen, B. L., Sumer, B. M., & Fredsoe, J., Turbulent oscilla<strong>to</strong>ry<br />

boundary lauers at high reynolds numbers, Journal of Fluid Mechanics,<br />

vol. 206, pp. 265-297, 1989.<br />

[27] Jones, W. P., & Launder, B. E., <strong>The</strong> prediction of laminarization with<br />

a two-equation model of turbulence, International Journal of Heat and<br />

Mass Transfer, vol. 15, pp. 301-314, 1972.<br />

[28] von Kármán, T., Mechanische Ähnlichkeit und Turbulenz, Proceedings<br />

3rd Int. Congr. <strong>Applied</strong> Mechanics, S<strong>to</strong>ckholm, pp. 85-105, 1930.<br />

[29] Kawamura, H., & Homma, K., DNS of an unsteady turbulent flow<br />

driven by a temporally sinusoidal pressure gradient, Turbulence and<br />

Shear <strong>Flow</strong> - 1, pp. 175-180, Begell House, New York, 1999.


BIBLIOGRAPHY 91<br />

[30] Kim J., Moin P., & Moser, R. D., Turbulence statistics in fully devel-<br />

oped channel flow at low Reynolds number, Journal of Fluid Mechan-<br />

ics, vol. 177, pp. 133-166, 1987.<br />

[31] Kolmogorov, A. N., Equations of turbulent motion of an incompressible<br />

turbulent fluid, Izv. Akad. Nauk SSSR Ser Phys., vol. 6, no. 1-2, p. 56,<br />

1942.<br />

[32] Labourasse, E. & Sagaut, P., Reconstruction of turbulent fluctua-<br />

tions using a hybrid RANS/LES approach, Journal of Computational<br />

Physics, vol. 182, pp. 301-336, 2002.<br />

[33] Launder, B. E. & Sharma, B. I., Application of the energy-dissipation<br />

model of turbulence <strong>to</strong> the calculation of flow near a spinning disk,<br />

Letters in Heat and Mass Transfer, vol. 1, pp. 131-137, 1974.<br />

[34] Launder, B. E. & Spalding, D. B., <strong>The</strong> numerical computation of turbu-<br />

lent flows, Computer Methods in <strong>Applied</strong> Mechanics and Engineering,<br />

vol. 3, pp. 269-289, 1974.<br />

[35] Lea, C. J., Second-moment closure computations of in-cylinder flows in<br />

idealised reciprocating engines, PhD <strong>The</strong>sis, University of Manchester,<br />

1994.<br />

[36] Lee, D.-Y., Park, S.-J., & Ro, S. T., Heat transfer by oscillating flow<br />

in a circular pipe with a sinusoidal wall temperature distribution, In-<br />

ternational Journal of Heat and Mass Transfer, vol. 38, pp. 2529-2537,<br />

1995.<br />

[37] Leonard, B. P., A stable and accurate convective modelling procedure<br />

based on quadratic upstream interpolation, Computer Methods in Ap-<br />

plied Mechanics and Engineering, vol. 19, pp. 59-98, 1979.


BIBLIOGRAPHY 92<br />

[38] Lodahl, C. R., Sumer, B. M., & Fredsoe, J., Turbulent combined os-<br />

cilla<strong>to</strong>ry flow and current in a pipe, Journal of Fluid Mechanics, vol.<br />

373, pp. 313-348, 1998.<br />

[39] Mansour, N. N., Kim, J., & Moin, P., Reynolds-stress and dissipation-<br />

rate budgets in a turbulent channel flow, Journal of Fluid Mechanics,<br />

vol. 194, pp. 15-44, 1988.<br />

[40] Menter, F. R., Influence of freestream values on k-ω turbulence model<br />

predictions, AIAA Journal, vol. 30, no. 6, pp. 1657-1659, 1992.<br />

[41] Menter, F. R., Improved two-equation k-ω turbulence models for aero-<br />

dynamic flows, NASA, no. TM-103975, 1992.<br />

[42] Menter, F. R., Two-equation eddy-viscosity turbulence models for en-<br />

gineering applications, AIAA Journal, vol. 32, pp. 1598-1605, 1994.<br />

[43] Monin, A. S. & Yaglom, A. M., Statistical Fluid Mechanics: Mechanics<br />

of Turbulence, vol. 2, MIT Press, 1975.<br />

[44] Moser, R. D., Kim, J., & Mansour, N. N., Direct numerical simulation<br />

of turbulent channel flow up <strong>to</strong> Retau=590, Physics of Fluids, vol. 11,<br />

pp. 943-945, 1999.<br />

[45] Nai<strong>to</strong>h, K. & Kuwahara, K., Large Eddy Simulation and direct simula-<br />

tion of compressible turbulence and combusting flows in engines based<br />

on the Biscales method, Fluid Dynamics Research, vol. 10, no. 4-6, pp.<br />

299-325, 1992.<br />

[46] Ng, K. H. & Spalding, D. B., Turbulence model for boundary layers<br />

near walls, Physics of Fluids, vol. 15, no. 1, pp. 20-30, 1972.


BIBLIOGRAPHY 93<br />

[47] Niederschulte, M. A., Adrian, R. J., & Hanratty, T. J., Measurements<br />

of turbulent flow in a channel at low Reynolds number, Experiments in<br />

Fluids, vol. 9, pp. 222-230, 1990.<br />

[48] Ohmi, M. & Iguchi, M., Critical Reynolds number in an oscillating pipe<br />

flow, Bulletin of the JSME, vol. 25, no. 200, pp. 165-172, 1982.<br />

[49] Ohmi, M., Iguchi, M. & Urahata, I., Transition <strong>to</strong> turbulence in a<br />

pulsatile pipe flow part 1: wave forms and distribution of pulsatile<br />

velocities near transition region, Bulletin of the JSME, vol. 25, no.<br />

200, pp. 182-189, 1982.<br />

[50] Ohmi, M., Iguchi, M. & Urahata, I., <strong>Flow</strong> patterns and frictional losses<br />

in an oscillating pipe flow, Bulletin of the JSME, vol. 25, no. 200, pp.<br />

536-543, 1982.<br />

[51] Patankar, S. V., Numerical Heat Transfer and Fluid <strong>Flow</strong>, Series in<br />

Computational Methods in Mechanics and <strong>The</strong>rmal Sciences, Hemi-<br />

sphere Publishing Corp., McGraw-Hill, 1980.<br />

[52] Peng, S.-H., Davidson, L., & Holmberg, S., A modified low-Reynolds<br />

number k-ω model for recirculating flows, Journal of Fluids Engineer-<br />

ing, vol. 119, pp. 867-875, 1997.<br />

[53] Pope, S. B., Turbulent <strong>Flow</strong>s, Cambridge University Press, Cambridge,<br />

UK, 2000.<br />

[54] Prandtl, L., Über die ausgebildete turbulenz, ZAMM, vol. 5, pp. 136-<br />

139, 1925.<br />

[55] Prandtl, L., Über ein neues formelsystem der ausgebildeten turbulenz,<br />

Nachr. Akad. Wiss. Göttingen, pp. 6-19, 1945.


BIBLIOGRAPHY 94<br />

[56] Quemere, P., Zonal multi-domain RANS/LES simulations of turbulent<br />

flows, International Journal for Numerical Methods in Fluids, vol. 40,<br />

pp. 903-925, 2002.<br />

[57] Ramaprian, B. R. & Tu, S. W., Fully developed periodic turbulent<br />

pipe flow part 2: the detailed structure of the flow, Journal of Fluid<br />

Mechanics, vol. 137, pp. 59-81, 1983.<br />

[58] Reichardt, H., Vollstandige darstellung der turbulenten<br />

geschwindigkeitsverteilung in glatten leitungen, Zeitschrift für<br />

Angewandte Mathematik und Mechanik, vol. 31, pp. 208-219, 1951.<br />

[59] Reynolds, O., On the dynamical theory of incompressible viscous fluids<br />

and determination of the criterion, Philosophical Transactions of the<br />

Royal Society of London, serries A, vol. 186, pp. 123-164, 1895.<br />

[60] Reynolds, W. C., modelling of fluid motion in engines: an introduc-<br />

<strong>to</strong>ry review, Combustion modelling in Reciprocating Engines, Plenum<br />

Press, New York, pp. 41-68, 1980.<br />

[61] Robinson, D. F., Harris, J. E., & Hassan, H. A., Unified turbulence clo-<br />

sure model for axisymmetric and planar free shear flows, AIAA Journal,<br />

vol. 33, no. 12, pp. 2324-2331, 1995.<br />

[62] Rodi, W. & Spalding, D. B., A two parameter model of turbulence and<br />

its application <strong>to</strong> free jets, Wärme und S<strong>to</strong>ffübertragung, vol. 3, p. 85,<br />

1970.<br />

[63] Rotta, J. C., Statistische theorie nichthomogener turbulenz, Zeitschrift<br />

für Physik, vol. 129, pp. 547-572, 1951.


BIBLIOGRAPHY 95<br />

[64] Rotta, J. C.,<br />

Über eine methode zur berechnung turbulenter scher-<br />

strömungen, Aerodynamische Versuchanstalt Göttingen, Rep. 69 A 14,<br />

1968.<br />

[65] Saffman, P. G., A model for inhomogeneous turbulent flow, Proceedings<br />

of the Royal Society of London, vol. A317, pp. 417-433, 1970.<br />

[66] Sarpkaya, T., Experimental determination of the critical Reynolds<br />

number for pulsating poiseuille flow Transactions of the ASME: Jour-<br />

nal of Basic Engineering, no. 66-FE-5, 1966.<br />

[67] Scotti, A. & Piomelli, U., Numerical simulation of pulsating turbulent<br />

channel flow, American Institute of Physics, vol. 13, no. 5, pp. 1367-<br />

1384, 2001.<br />

[68] Shemer, L., Wygnanski, I., & Kit, E., Pulsating flow in a pipe, Journal<br />

of Fluid Mechanics, vol. 153, pp. 313-337, 1985.<br />

[69] Shima, N., Calculation of a variety of boundary layers with a second-<br />

moment closure applicable up <strong>to</strong> a wall, 7th Symposium on Turbulent<br />

Shear <strong>Flow</strong>s, Stanford University, Paper 5-3, 1989.<br />

[70] Siginer, A., Oscillating flow of a simple fluid in a pipe, International<br />

Journal of Engineering Science, vol. 29, no. 12, pp. 1557-1567, 1991.<br />

[71] Sleath, J. F. A., Turbulent oscilla<strong>to</strong>ry flow over rough beds, Journal of<br />

Fluid Mechanics, vol. 182, pp. 369-409, 1987.<br />

[72] Smith, B. R., <strong>The</strong> k-kl turbulence model and wall layer model for<br />

compressible flows, AIAA Paper, no. 90-1483, 1990.<br />

[73] Spalart, P. R., A one-equation turbulence model for aerodynamic flows,<br />

AIAA Paper, no. 92-439, 1992.


BIBLIOGRAPHY 96<br />

[74] Spalding, D. B., <strong>The</strong> calculation of the length scale of turbulence in<br />

some turbulent boundary layers remote from walls, Imperial College,<br />

Heat Transfer Section Report, no. TWF/TN/31, 1967.<br />

[75] Spalding, D.B., <strong>The</strong> prediction of two-dimensional steady turbulent<br />

flows, Imperial College, Heat and Mass Transfer Section Report, no.<br />

EF/TN/A/16, 1969.<br />

[76] Speziale, C. G., Abid, R., & Anderson, E. C., A critical evaluation<br />

of two-equation models for near wall turbulence, AIAA Paper, no. 90-<br />

1481, 1990.<br />

[77] Suga, K., Development and application of a non-linear eddy viscosity<br />

model sensitized <strong>to</strong> stress and strain invariants, PhD thesis, University<br />

of Manchester Institute of Science and Technology (<strong>UMIST</strong>), 1995.<br />

[78] Tabaczynski, R. J., Turbulent flows in reciprocating internal combus-<br />

tion engines, Internal Combustion Engineering: Science and Technol-<br />

ogy, pp. 243-285, 1990.<br />

[79] Tardu, S. F., Binder, G., & Blackwelder, R. F., Turbulent channel flow<br />

with large-amplitude velocity oscillations, Journal of Fluid Mechanics,<br />

vol. 267, pp. 109-151, 1994.<br />

[80] Taylor, G. I., Eddy motion in the atmosphere, Philosophical Transac-<br />

tions of the Royal Society of London, serries A, p. 215, 1915.<br />

[81] Townsend, A. A., <strong>The</strong> Structure of Turbulent Shear <strong>Flow</strong>, Cambridge<br />

University Press, 1956.


BIBLIOGRAPHY 97<br />

[82] Tu, S. W. & Ramaprian, B. R., Fully developed periodic turbulent<br />

pipe flow part 1: main experimental results and comparison with pre-<br />

dictions, Journal of Fluid Mechanics, vol. 137, pp. 31-58, 1983.<br />

[83] Versteeg, H. K. & Malalasekera, W., An Introduction <strong>to</strong> Computational<br />

Fluid Dynamics: <strong>The</strong> Finite Volume Method, Pearson Education, Ltd.,<br />

Harlow, UK, 1995.<br />

[84] Viegas, J. R. & Rubesin, M. W., <strong>Wall</strong>-function boundary conditions in<br />

the solution of the Navier-S<strong>to</strong>kes equations for complex compressible<br />

flows, AIAA, 16th Fluid & Plasma Dynamics Conference, MA, USA,<br />

1983.<br />

[85] Walther, C., Kühl, H.-D., & Schulz, S., Numerical investigations on<br />

the heat transfer in turbulent oscillating pipe flow, Journal of Heat<br />

and Mass Transfer, vol. 36, pp. 135-141, 2000.<br />

[86] Watkins, A. P., Calculation of flow and heat transfer in the combustion<br />

chamber of a reciprocating engine, MSc <strong>The</strong>sis, Imperial College of<br />

Science, Technology, & Medicine, University of London, 1973.<br />

[87] Wei, T. & Willmarth, W. W., Reynolds number effects on the structure<br />

of a turbulent channel flow, Journal of Fluid Mechanics, vol. 204, pp.<br />

57-95, 1989.<br />

[88] Wilcox, D. C., Rubesin, W. M., Progress in turbulence modelling for<br />

complex flow fields including effects of compressibility, NASA Technical<br />

Paper, no. 1517, 1980.<br />

[89] Wilcox, D. C., Reassessment of the scale-determining equation for ad-<br />

vanced turbulence models, AIAA Journal, vol. 26, no. 11, pp. 1299-<br />

1310, 1988.


BIBLIOGRAPHY 98<br />

[90] Wilcox, D. C., Two-equation eddy-viscosity turbulence models for en-<br />

gineering applications, AIAA Journal, vol. 32, no. 8, pp. 1598-1605,<br />

1994.<br />

[91] Wilcox, D. C., Turbulence modelling for CFD, 2nd ed., DCW Indus-<br />

tries, La Cañada, California, 1998.<br />

[92] Yap, C. R., Turbulent heat and momentum transfer in recirculating<br />

and impinging flows, PhD <strong>The</strong>sis, University of Manchester, 1987.


Figures<br />

99


FIGURES 100<br />

U +<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

10 0<br />

10 1<br />

+<br />

+<br />

+<br />

+<br />

Reichardt’s law<br />

DNS: Reτ = 180 [30]<br />

DNS: Reτ = 395 [44]<br />

DNS: Reτ = 584 [44]<br />

◦ Exp.: Reτ = 708 [87]<br />

▽ Exp.: Reτ = 921 [47]<br />

+<br />

10 2<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

x +<br />

x<br />

x xxx<br />

x<br />

x<br />

x +<br />

x +<br />

+<br />

10 3<br />

y +<br />

10 4<br />

Exp.: Reτ = 1017 [47]<br />

+ Exp.: Reτ = 1655 [47]<br />

⋄ Exp.: Reτ = 2340 [11]<br />

× Exp.: Reτ = 4800 [11]<br />

△ Exp.: Reτ = 8150 [11]<br />

Figure 3.1: Reichardt’s law <strong>to</strong> estimate U +


FIGURES 101<br />

− 〈uv〉 +<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

-0.1<br />

+ + +<br />

+ +<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

10 1<br />

Reichardt: Reτ = 180<br />

Reichardt: Reτ = 584<br />

Reichardt: Reτ = 1017<br />

Reichardt: Reτ = 1655<br />

+<br />

+<br />

x<br />

x<br />

x<br />

+<br />

+<br />

+<br />

++++++++++++++++++++++++++++++++<br />

x x x x<br />

+++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++<br />

x<br />

x<br />

x<br />

x<br />

x<br />

10 2<br />

x<br />

x<br />

y +<br />

10 3<br />

DNS: Reτ = 180 [30]<br />

+ DNS: Reτ = 584 [44]<br />

⋄ Exp.: Reτ = 1017 [47]<br />

× Exp.: Reτ = 1655 [47]<br />

Figure 3.2: Reichardt’s law applied <strong>to</strong> − 〈uv〉 +


FIGURES 102<br />

− 〈uv〉 +<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

-0.1<br />

+ + +<br />

+ +<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

10 1<br />

Profile: Reτ = 180<br />

Profile: Reτ = 584<br />

Profile: Reτ = 1017<br />

Profile: Reτ = 1655<br />

+<br />

+<br />

x<br />

x<br />

x<br />

+<br />

+<br />

+<br />

++++++++++++++++++++++++++++++++<br />

x x x x<br />

+++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++<br />

x<br />

x<br />

x<br />

x<br />

x<br />

10 2<br />

x<br />

x<br />

y +<br />

10 3<br />

DNS: Reτ = 180 [30]<br />

+ DNS: Reτ = 584 [44]<br />

⋄ Exp.: Reτ = 1017 [47]<br />

× Exp.: Reτ = 1655 [47]<br />

Figure 3.3: A revised profile for − 〈uv〉 +


FIGURES 103<br />

k +<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

10 0<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

10 1<br />

Profile: Reτ = 180<br />

Profile: Reτ = 395<br />

Profile: Reτ = 584<br />

+<br />

+<br />

+ + +<br />

+++++++++++++++++++++++++++++++++ +++++++++++++++++++++++++++<br />

+<br />

10 2<br />

Figure 3.4: A profile for k +<br />

10 3<br />

y +<br />

DNS: Reτ = 180 [30]<br />

+ DNS: Reτ = 395 [44]<br />

⋄ DNS: Reτ = 584 [44]


FIGURES 104<br />

〈uu〉 9<br />

+<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

10 0<br />

+<br />

10 1<br />

Profile: Reτ = 180<br />

Profile: Reτ = 395<br />

Profile: Reτ = 584<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

10 2<br />

+<br />

+<br />

+<br />

+<br />

+<br />

10 3<br />

y +<br />

DNS: Reτ = 180 [30]<br />

+ DNS: Reτ = 395 [44]<br />

⋄ DNS: Reτ = 584 [44]<br />

Figure 3.5: A profile for 〈uu〉 +


FIGURES 105<br />

1.4<br />

〈vv〉 +<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

10 0<br />

+<br />

10 1<br />

Profile: Reτ = 180<br />

Profile: Reτ = 395<br />

Profile: Reτ = 584<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+ +<br />

+<br />

10 2<br />

+<br />

+<br />

+<br />

+<br />

+<br />

10 3<br />

y +<br />

DNS: Reτ = 180 [30]<br />

+ DNS: Reτ = 395 [44]<br />

⋄ DNS: Reτ = 584 [44]<br />

Figure 3.6: A profile for 〈vv〉 +


FIGURES 106<br />

0.20<br />

k +<br />

“<br />

y +” 2<br />

0.15<br />

0.10<br />

0.05<br />

0.20<br />

〈uu〉 +<br />

“<br />

y +” 2<br />

0.00<br />

0 2 4 6 8 10<br />

0.15<br />

0.10<br />

0.05<br />

y +<br />

0.00<br />

0 2 4 6 8 10<br />

y +<br />

−〈uv〉 +<br />

“<br />

y +” 3<br />

〈vv〉 +<br />

“<br />

y +” 4<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

0.0000<br />

0 2 4 6 8 10<br />

0.00020<br />

0.00015<br />

0.00010<br />

0.00005<br />

y +<br />

0.00000<br />

0 2 4 6 8 10<br />

▽ DNS: Reτ = 180 [30]<br />

⋄ DNS: Reτ = 395 [44]<br />

△ DNS: Reτ = 584 [44]<br />

Figure 3.7: <strong>Near</strong>-wall behaviour of flow parameters<br />

y +


FIGURES 107<br />

y ∗<br />

y ∗<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0 100 200 300 400 500 600<br />

100<br />

80<br />

60<br />

40<br />

20<br />

y +<br />

0<br />

0 20 40 60 80 100<br />

y +<br />

▽ DNS: Reτ = 180 [30]<br />

⋄ DNS: Reτ = 395 [44]<br />

△ DNS: Reτ = 584 [44]<br />

y ∗ = y +<br />

Figure 3.8: y ∗ vs. y +


FIGURES 108<br />

U ∗<br />

or<br />

U +<br />

25<br />

20<br />

15<br />

10<br />

5<br />

++++++++++++++++++++++++++++++++++++++++++++<br />

+++++++++++++++++++++++++++++++++++++++++<br />

+++<br />

xxxx xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxxxxxxx x xxxxxxxx<br />

0 + x<br />

0 100 200 300 400 500 600<br />

y∗ or y +<br />

U + vs. y + :<br />

▽ DNS: Reτ = 180 [30]<br />

⋄ DNS: Reτ = 395 [44]<br />

△ DNS: Reτ = 584 [44]<br />

U ∗ vs. y ∗ :<br />

✄✂ ✁ DNS: Reτ = 180 [30]<br />

+ DNS: Reτ = 395 [44]<br />

x DNS: Reτ = 584 [44]<br />

Figure 3.9: U ∗ superimposed on U +


FIGURES 109<br />

y ∗ v2<br />

y ∗ v2<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0 100 200 300 400 500 600<br />

100<br />

80<br />

60<br />

40<br />

20<br />

y +<br />

0<br />

0 20 40 60 80 100<br />

y +<br />

▽ DNS: Reτ = 180 [30]<br />

⋄ DNS: Reτ = 395 [44]<br />

△ DNS: Reτ = 584 [44]<br />

y ∗ v2 = y +<br />

Figure 3.10: y ∗ v2 vs. y +


FIGURES 110<br />

U ∗ v2<br />

or<br />

U +<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

x<br />

+<br />

x<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

++++++++++++++++++++++++++++++++ +++++++++++++++++++++++++++++++++++++++++++++++<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx<br />

5<br />

0 + x<br />

0 100 200 300 400 500 600<br />

y∗ v2 or y +<br />

U + vs. y + :<br />

▽ DNS: Reτ = 180 [30]<br />

⋄ DNS: Reτ = 395 [44]<br />

△ DNS: Reτ = 584 [44]<br />

U ∗ v2 vs. y ∗ v2:<br />

✄✂ ✁ DNS: Reτ = 180 [30]<br />

+ DNS: Reτ = 395 [44]<br />

x DNS: Reτ = 584 [44]<br />

Figure 3.11: U ∗ v2 superimposed on U +


FIGURES 111<br />

〈U〉 +<br />

〈U〉 +<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.25 0.5 0.75 1<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

a<br />

y/δ<br />

0<br />

0 0.25 0.5 0.75 1<br />

c<br />

y/δ<br />

〈U〉 +<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.25 0.5 0.75 1<br />

a {<br />

b{<br />

c {<br />

Figure 5.1: 〈U〉 + vs y/δ in the steady flow case<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

b<br />

y/δ<br />

k-ε with log law<br />

Subgrid k-ε<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

✄✂ ✁ DNS [30]<br />

Reichardt’s law [58]


FIGURES 112<br />

k +<br />

k +<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 0.25 0.5 0.75 1<br />

5<br />

4<br />

3<br />

2<br />

1<br />

a<br />

y/δ<br />

0<br />

0 0.25 0.5 0.75 1<br />

c<br />

y/δ<br />

k +<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 0.25 0.5 0.75 1<br />

a {<br />

b{<br />

c {<br />

Figure 5.2: k + vs y/δ in the steady flow case<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

b<br />

y/δ<br />

k-ε with log law<br />

Subgrid k-ε<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

✄✂ ✁ DNS [30]<br />

analytical profile


FIGURES 113<br />

〈U〉 +<br />

〈U〉 +<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

10 0 0<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

10 0 0<br />

10 1<br />

10 1<br />

c<br />

a<br />

10 2<br />

10 2<br />

y +<br />

y +<br />

.<br />

〈U〉 +<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

10 0 0<br />

a {<br />

b{<br />

c {<br />

Figure 5.3: 〈U〉 + vs y + in the steady flow case<br />

10 1<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

b<br />

10 2<br />

k-ε with log law<br />

Subgrid k-ε<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

y +<br />

✄✂ ✁ DNS [30]<br />

Reichardt’s law [58]


FIGURES 114<br />

k +<br />

k +<br />

5<br />

4<br />

3<br />

2<br />

1<br />

10 0 0<br />

5<br />

4<br />

3<br />

2<br />

1<br />

10 0 0<br />

10 1<br />

10 1<br />

a<br />

c<br />

10 2<br />

10 2<br />

y +<br />

y +<br />

k +<br />

5<br />

4<br />

3<br />

2<br />

1<br />

10 0 0<br />

a {<br />

b{<br />

c {<br />

Figure 5.4: k + vs y + in the steady flow case<br />

10 1<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

b<br />

10 2<br />

y +<br />

k-ε with log law<br />

Subgrid k-ε<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

✄✂ ✁ DNS [30]<br />

analytical profile


FIGURES 115<br />

U<br />

U ss<br />

U<br />

U ss<br />

3.0 3.0<br />

2.5 2.5<br />

2.0 2.0<br />

1.5 1.5<br />

1.0 1.0<br />

0.5 0.5<br />

0.0 0.0<br />

-0.5 -0.5<br />

-1.0 -1.0<br />

-1.5 -1.5<br />

-2.0 -2.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a<br />

period<br />

3.0 3.0<br />

2.5 2.5<br />

2.0 2.0<br />

1.5 1.5<br />

1.0 1.0<br />

0.5 0.5<br />

0.0 0.0<br />

-0.5 -0.5<br />

-1.0 -1.0<br />

-1.5 -1.5<br />

-2.0 -2.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a {<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

b{<br />

✄✂ ✁ DNS [29] prescribed<br />

b<br />

period<br />

k-ε log law<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

∂〈P 〉<br />

∂x<br />

( ∂〈P 〉<br />

∂x )<br />

( ∂〈P 〉<br />

∂x ) ss<br />

( ∂〈P 〉<br />

( ∂〈P 〉<br />

∂x ) ss<br />

Figure 5.5: Bulk flow variation in the periodic pressure case<br />

∂x )<br />

× 10 −1<br />

× 10 −1


FIGURES 116<br />

τw<br />

(τw) ss<br />

τw<br />

(τw) ss<br />

3.0 3.0<br />

2.5 2.5<br />

2.0 2.0<br />

1.5 1.5<br />

1.0 1.0<br />

0.5 0.5<br />

0.0 0.0<br />

-0.5 -0.5<br />

-1.0 -1.0<br />

-1.5 -1.5<br />

-2.0 -2.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a<br />

period<br />

3.0 3.0<br />

2.5 2.5<br />

2.0 2.0<br />

1.5 1.5<br />

1.0 1.0<br />

0.5 0.5<br />

0.0 0.0<br />

-0.5 -0.5<br />

-1.0 -1.0<br />

-1.5 -1.5<br />

-2.0 -2.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a {<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

b{<br />

✄✂ ✁ DNS [29] prescribed<br />

b<br />

period<br />

k-ε log law<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

∂〈P 〉<br />

∂x<br />

( ∂〈P 〉<br />

( ∂〈P 〉<br />

∂x ) ss<br />

∂x )<br />

( ∂〈P 〉<br />

( ∂〈P 〉<br />

∂x ) ss<br />

∂x )<br />

Figure 5.6: <strong>Wall</strong> shear stress variation in the periodic pressure case<br />

× 10 −1<br />

× 10 −1


FIGURES 117<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

phase angle<br />

Subgrid k-ω<br />

✄✂ ✁ DNS [29]<br />

Figure 5.7: Variables with time at y/δ = 0.1 (prescribed<br />

∂〈P 〉<br />

∂x )


FIGURES 118<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

phase angle<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.8: Variables with time at y/δ = 0.2 (prescribed<br />

∂〈P 〉<br />

∂x )


FIGURES 119<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

phase angle<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.9: Variables with time at y/δ = 0.5 (prescribed<br />

∂〈P 〉<br />

∂x )


FIGURES 120<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

phase angle<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.10: Variables with time at y/δ = 0.9 (prescribed<br />

∂〈P 〉<br />

∂x )


FIGURES 121<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = 0<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

y/δ<br />

phase angle = π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.11: 〈U〉 vs y/δ snapshots through time (prescribed<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 3π<br />

4<br />

∂〈P 〉<br />

) - part 1<br />

∂x<br />

y/δ


FIGURES 122<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = π<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

y/δ<br />

phase angle = 3π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = 5π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.12: 〈U〉 vs y/δ snapshots through time (prescribed<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 7π<br />

4<br />

∂〈P 〉<br />

) - part 2<br />

∂x<br />

y/δ


FIGURES 123<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

phase angle = 0<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

y/δ<br />

phase angle = π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.13: k vs y/δ snapshots through time (prescribed<br />

5<br />

phase angle = π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 3π<br />

4<br />

∂〈P 〉<br />

) - part 1<br />

∂x<br />

y/δ


FIGURES 124<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

phase angle = π<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

y/δ<br />

phase angle = 3π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.14: k vs y/δ snapshots through time (prescribed<br />

5<br />

phase angle = 5π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 7π<br />

4<br />

∂〈P 〉<br />

) - part 2<br />

∂x<br />

y/δ


FIGURES 125<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

phase angle = 0<br />

10 1<br />

10 2<br />

y +<br />

phase angle = π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.15: 〈U〉 vs y + snapshots through time (prescribed<br />

phase angle = π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 3π<br />

4<br />

∂〈P 〉<br />

) - part 1<br />

∂x<br />

y +


FIGURES 126<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

phase angle = π<br />

10 1<br />

10 2<br />

y +<br />

phase angle = 3π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.16: 〈U〉 vs y + snapshots through time (prescribed<br />

phase angle = 5π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 7π<br />

4<br />

∂〈P 〉<br />

) - part 2<br />

∂x<br />

y +


FIGURES 127<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

5<br />

10 0<br />

0<br />

phase angle = 0<br />

10 1<br />

10 2<br />

y +<br />

phase angle = π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.17: k vs y + snapshots through time (prescribed<br />

5<br />

phase angle = π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 3π<br />

4<br />

∂〈P 〉<br />

) - part 1<br />

∂x<br />

y +


FIGURES 128<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

5<br />

10 0<br />

0<br />

phase angle = π<br />

10 1<br />

10 2<br />

y +<br />

phase angle = 3π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.18: k vs y + snapshots through time (prescribed<br />

5<br />

phase angle = 5π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 7π<br />

4<br />

∂〈P 〉<br />

) - part 2<br />

∂x<br />

y +


FIGURES 129<br />

( ∂〈P 〉<br />

∂x )<br />

( ∂〈P 〉<br />

∂x ) ss<br />

( ∂〈P 〉<br />

∂x )<br />

( ∂〈P 〉<br />

∂x ) ss<br />

25.0 25.0<br />

20.0 20.0<br />

15.0 15.0<br />

10.0 10.0<br />

5.0 5.0<br />

0.0 0.0<br />

-5.0 -5.0<br />

-10.0 -10.0<br />

-15.0 -15.0<br />

-20.0 -20.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a<br />

period<br />

25.0 25.0<br />

20.0 20.0<br />

15.0 15.0<br />

10.0 10.0<br />

5.0 5.0<br />

0.0 0.0<br />

-5.0 -5.0<br />

-10.0 -10.0<br />

-15.0 -15.0<br />

-20.0 -20.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a {<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

DNS<br />

∂〈P 〉<br />

∂x [29]<br />

b{<br />

b<br />

period<br />

k-ε log law<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

Figure 5.19: Pressure variation in the periodic bulk flow


FIGURES 130<br />

τw<br />

(τw) ss<br />

τw<br />

(τw) ss<br />

3.0 3.0<br />

2.5 2.5<br />

2.0 2.0<br />

1.5 1.5<br />

1.0 1.0<br />

0.5 0.5<br />

0.0 0.0<br />

-0.5 -0.5<br />

-1.0 -1.0<br />

-1.5 -1.5<br />

-2.0 -2.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a<br />

period<br />

3.0 3.0<br />

2.5 2.5<br />

2.0 2.0<br />

1.5 1.5<br />

1.0 1.0<br />

0.5 0.5<br />

0.0 0.0<br />

-0.5 -0.5<br />

-1.0 -1.0<br />

-1.5 -1.5<br />

-2.0 -2.0<br />

0.0 0.5 1.0 1.5 2.0<br />

a {<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

b{<br />

✄✂ ✁ DNS τw [29] DNS<br />

b<br />

period<br />

k-ε log law<br />

Subgrid k-ε<br />

Subgrid k-ω<br />

∂〈P 〉<br />

∂x [29]<br />

( ∂〈P 〉<br />

( ∂〈P 〉<br />

∂x ) ss<br />

∂x )<br />

( ∂〈P 〉<br />

( ∂〈P 〉<br />

∂x ) ss<br />

∂x )<br />

Figure 5.20: <strong>Wall</strong> shear stress variation in the periodic bulk flow case<br />

× 10 −1<br />

× 10 −1


FIGURES 131<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

phase angle<br />

Subgrid k-ω<br />

✄✂ ✁ DNS [29]<br />

Figure 5.21: Variables with time at y/δ = 0.1 (prescribed U)


FIGURES 132<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

phase angle<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.22: Variables with time at y/δ = 0.2 (prescribed U)


FIGURES 133<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

phase angle<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.23: Variables with time at y/δ = 0.5 (prescribed U)


FIGURES 134<br />

〈U〉<br />

(Uτ ) ss<br />

k<br />

(U 2 τ ) ss<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 π/2 π 3π/2 2π<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

phase angle<br />

0<br />

0 π/2 π 3π/2 2π<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

phase angle<br />

Subgrid k-ω<br />

k-ε log law<br />

Figure 5.24: Variables with time at y/δ = 0.9 (prescribed U)


FIGURES 135<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = 0<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

y/δ<br />

phase angle = π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

Subgrid k-ω<br />

k-ε log law<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 3π<br />

4<br />

Figure 5.25: 〈U〉 vs y/δ snapshots through time (prescribed U) - part 1<br />

y/δ


FIGURES 136<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = π<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

y/δ<br />

phase angle = 3π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

phase angle = 5π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

30<br />

20<br />

10<br />

Subgrid k-ω<br />

k-ε log law<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 7π<br />

4<br />

Figure 5.26: 〈U〉 vs y/δ snapshots through time (prescribed U) - part 2<br />

y/δ


FIGURES 137<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

phase angle = 0<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

y/δ<br />

phase angle = π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

Subgrid k-ω<br />

k-ε log law<br />

5<br />

phase angle = π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 3π<br />

4<br />

Figure 5.27: k vs y/δ snapshots through time (prescribed U) - part 1<br />

y/δ


FIGURES 138<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

phase angle = π<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

y/δ<br />

phase angle = 3π<br />

2<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

Subgrid k-ω<br />

k-ε log law<br />

5<br />

phase angle = 5π<br />

4<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

5<br />

0<br />

0.00 0.20 0.40 0.60 0.80<br />

y/δ<br />

phase angle = 7π<br />

4<br />

Figure 5.28: k vs y/δ snapshots through time (prescribed U) - part 2<br />

y/δ


FIGURES 139<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

phase angle = 0<br />

10 1<br />

10 2<br />

y +<br />

phase angle = π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

phase angle = π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 3π<br />

4<br />

Figure 5.29: 〈U〉 vs y + snapshots through time (prescribed U) - part 1<br />

y +


FIGURES 140<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

phase angle = π<br />

10 1<br />

10 2<br />

y +<br />

phase angle = 3π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

〈U〉<br />

(Uτ ) ss<br />

〈U〉<br />

(Uτ ) ss<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

30<br />

20<br />

10<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

phase angle = 5π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 7π<br />

4<br />

Figure 5.30: 〈U〉 vs y + snapshots through time (prescribed U) - part 2<br />

y +


FIGURES 141<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

5<br />

10 0<br />

0<br />

phase angle = 0<br />

10 1<br />

10 2<br />

y +<br />

phase angle = π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

5<br />

phase angle = π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 3π<br />

4<br />

Figure 5.31: k vs y + snapshots through time (prescribed U) - part 1<br />

y +


FIGURES 142<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

5<br />

10 0<br />

0<br />

phase angle = π<br />

10 1<br />

10 2<br />

y +<br />

phase angle = 3π<br />

2<br />

10 1<br />

10 2<br />

y +<br />

Low-Re k-ε<br />

Subgrid k-ε<br />

✄✂ ✁ DNS [29]<br />

k<br />

(U 2 τ ) ss<br />

k<br />

(U 2 τ ) ss<br />

5<br />

10 0<br />

0<br />

10 0<br />

0<br />

Subgrid k-ω<br />

k-ε log law<br />

5<br />

phase angle = 5π<br />

4<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

y +<br />

phase angle = 7π<br />

4<br />

Figure 5.32: k vs y + snapshots through time (prescribed U) - part 2<br />

y +

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!