You are on page 1of 72

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

net/publication/318562275

Turbulence Modeling - A Review

Technical Report · November 2018


DOI: 10.13140/RG.2.2.35857.33129/2

CITATIONS READS
0 1,769

1 author:

Ideen Sadrehaghighi
CFD Open Series
23 PUBLICATIONS   24 CITATIONS   

SEE PROFILE

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

Independent CFD Reasercher View project

All content following this page was uploaded by Ideen Sadrehaghighi on 16 November 2018.

The user has requested enhancement of the downloaded file.


1

CFD Open Series


Revision 1.85.5

Turbulence Modeling
A Review
Ideen Sadrehaghighi, Ph.D.

A sketch by Leonardo da Vinci


of turbulent flow caused by
pouring water into a pool

ANNAPOLIS, MD
2

Contents

1 Introduction .................................................................................................................................. 7
1.1 Background............................................................................................................................................................... 7
1.2 Physical Description ............................................................................................................................................. 8

2 Turbulence Essentials ............................................................................................................. 10


2.1 Physical Perspectives ........................................................................................................................................ 10
2.2 Components Attributing to Complexity of Physics in Turbulence ................................................. 10
Enhanced Diffusion ................................................................................................................................ 11
Stability Criteria ...................................................................................................................................... 11
Time Dependency ................................................................................................................................... 11
Eddies and Spectral Length ................................................................................................................ 12
Non-Linearity Effects ............................................................................................................................ 12
Separation and Drag Reduction ........................................................................................................ 13
2.3 Remarks Regarding Transition Phenomenon ......................................................................................... 13
Transition Types ..................................................................................................................................... 15
2.3.1.1 Natural Transition.................................................................................................... 15
2.3.1.2 Bypass Transition..................................................................................................... 15
2.3.1.3 Reverse Transition ................................................................................................... 16
2.3.1.4 Separated Flow Transition....................................................................................... 16
2.3.1.5 Effect of Geometry .................................................................................................. 16
2.3.1.6 Boundary Heat Transfer .......................................................................................... 16
What Triggers Transition? .................................................................................................................. 17

3 Reynolds Averaging and RANS Modeling (Linear Modeling) .................................... 18


3.1 Time Averaging .................................................................................................................................................... 18
Un-Steady RANS (URANS)................................................................................................................... 19
3.2 Reynolds Averaged Naiver-Stokes (RANS) .............................................................................................. 19
Comments on Reynolds Averaged Naiver-Stokes (RANS) ..................................................... 20
3.3 Closure Problem (Boussinesq Assumption) ............................................................................................ 21
3.4 Turbulence Models ............................................................................................................................................. 21
Turbulent Viscosity, Velocity, and Length Scale ........................................................................ 23
3.4.1.1 Eddy Viscosity (RANS) Models ................................................................................. 23
3.4.1.2 Algebraic Model (Zero Equation) - Mixing Length .................................................. 24
3.4.1.3 Baldwin-Lomax ........................................................................................................ 24
One-Equation Model .............................................................................................................................. 24
3.4.2.1 k - Model.................................................................................................................. 25
3.4.2.2 Spallart - Allmaras.................................................................................................... 25
Two Equation Model ............................................................................................................................. 25
3.4.3.1 Standard Transport Equation for k .......................................................................... 26
3.4.3.2 Standard Transport Equation for ε .......................................................................... 27
3.4.3.3 The RNG Model ....................................................................................................... 28
3.4.3.4 The Realizable κ-ε Model ........................................................................................ 28
3.4.3.5 Issues Relating to κ-ε Turbulence Modeling............................................................ 28
3.4.3.6 The κ-ω and SST Models .......................................................................................... 29
3.4.3.7 Wilcox (Standard) κ-ω Model .................................................................................. 30
3

3.4.3.8 The Baseline (BSL) κ-ω Model ................................................................................. 30


3.4.3.9 SST κ-ω Model ......................................................................................................... 30
Case Study 1 – Subsonic Flow for Turbulence Models of (NACA)-0012 Airfoil ............ 32
3.4.4.1 Computational Method ........................................................................................... 32
3.4.4.2 Results and Discussion ............................................................................................ 33
3.4.4.3 Conclusions.............................................................................................................. 35
3.5 Near Wall Turbulence Modeling ................................................................................................................... 36
Modeling Flow Near the Wall............................................................................................................. 37
3.5.1.1 Wall Function Method............................................................................................. 37
3.5.1.2 Low-Reynolds-Number Method .............................................................................. 37
Case Study 2 - Wall-Modelling Strategies in Large Eddy Simulation of Separated
High- Reynolds-Number Flows ........................................................................................................................... 38
Case Study 3 - Numerical Modeling of Turbulence in Turbine Stages .............................. 39
3.5.3.1 Background and Literature Survey .......................................................................... 39
3.5.3.2 Transition Phenomenon as Relates to Gas Turbines ............................................... 40
3.5.3.3 Example of Modeling of Turbine Stage with Twisted Rotor Blade with Different
Turbulence Models ...................................................................................................................... 42
3.5.3.4 Results ..................................................................................................................... 43

4 Beyond the Boussinesq Approximation (Non-Linear Modeling) ............................. 45


4.1 Boussinesq-Approximation Deficiencies .................................................................................................. 45
4.2 Nonlinear Constitutive Relations ................................................................................................................. 45
Case Study – 3D Simulation of Flow Past a Cylinder using Nonlinear Turbulence
Model 46
4.2.1.1 Literature Survey ..................................................................................................... 47
4.2.1.2 Basic Equations ........................................................................................................ 47
4.2.1.3 Numerical Strategies ............................................................................................... 48
4.2.1.4 Geometry and Meshing ........................................................................................... 49
4.2.1.5 Results and Discussion ............................................................................................ 49
4.2.1.6 Conclusions.............................................................................................................. 50
4.3 2 Order Closures (Reynolds Stress Models - RSM) ............................................................................ 51
nd

4.4 LES Model ............................................................................................................................................................... 51


Filter Definition ....................................................................................................................................... 52
Filtered Incompressible N-S Equations ......................................................................................... 52
Numerical Methods for LES ................................................................................................................ 54
Implicit vs Explicit Filtering ............................................................................................................... 54
Quantitative Aspects of Comparison of RANS vs. LES Models ............................................. 54
Motivations for Coupling Methods Between RANS and LES ................................................. 55
4.4.6.1 Principal Approaches to Coupling LES with RANS ................................................... 56
4.4.6.2 Results and Discussion for Segregated Modeling ................................................... 57
4.4.6.3 Coupling in RANS/LES Interface............................................................................... 58
4.5 DES Model .............................................................................................................................................................. 59
4.6 DNS Model .............................................................................................................................................................. 60
The Reynolds Number Constraint ................................................................................................... 61
Numerical Considerations ................................................................................................................... 61
4.6.2.1 Spectral Methods .................................................................................................... 62
4.6.2.2 Finite Differencing ................................................................................................... 62
4.7 Strategies for Turbulence Modelling........................................................................................................... 63
4

Physical Aspects (Modification to Simple RANS Models) ...................................................... 64


Complex RANS Models.......................................................................................................................... 64
Role of Grid Refinement ....................................................................................................................... 65
Outlook ........................................................................................................................................................ 65
4.8 Modeling To Choose From............................................................................................................................... 66

5 Classification of Turbulence Models .................................................................................. 68


5.1 Comparisons of various Turbulence Models ........................................................................................... 69
5.2 Numerical Considerations ............................................................................................................................... 70
Elementary Time-Marching Methods............................................................................................. 70
5.2.1.1 Block-Implicit Methods ........................................................................................... 70

List of Tables:
Table 3.1 Constant values for κ-ε turbulence model equations ............................................................... 27
Table 3.2 Constant values for SST κ-ω turbulence model equations..................................................... 31
Table 5.1 Advantages & Disadvantages of Different Turbulence Models ............................................ 69

List of Figures:
Figure 1.1 A sketch of turbulence by Leonardo Da Vinci, picture taken from (Bingham) ............... 8
Figure 2.1 Flow Pattern Pass a Circle.................................................................................................................. 10
Figure 2.2 Large Eddies in a Turbulent Boundary Layer ............................................................................ 12
Figure 2.3 Velocity Profile for Laminar vs Turbulent Flows ..................................................................... 13
Figure 2.4 Turbulence Transition Region ......................................................................................................... 13
Figure 2.5 Snapshots of the Stream-Wise Velocity Component for (A) Adiabatic Wall , (B)
Heated Wall, and (C) Cooled Wall ............................................................................................................................. 14
Figure 3.1 Relationship between averaged variables; a) Steady flow b) Un-Steady flow ...... 19
Figure 3.2 Hierchey of Turbulence Models based on Relative Importance of Numerics vs.
Computational Costs - (Courtsy of Tenzor) .......................................................................................................... 22
Figure 3.3 Hierarchy of Turbulence Models in General .............................................................................. 22
Figure 3.4 Lift Coefficient at Stall (AoA) against Number of Cells........................................................... 32
Figure 3.5 Comparison Between Experimental Data [Abbott et al.] and Three Different
Turbulent Models Simulation Results for Lift Coefficient in NACA 0012 Airfoil .................................. 33
Figure 3.6 Comparison Between Experimental Data for Transitional Boundary Layer and
Different Turbulent Models on the Drag Coefficient of NACA-0012 Airfoil ........................................... 34
Figure 3.7 Contours of velocity magnitude at 9° (Top) and 16° (Bottom) AoA with the Spalart-
Allmaras turbulence model ......................................................................................................................................... 35
Figure 3.8 Zones in Turbulent B. L. for a typical Incompressible flow over a smooth flat plate 36
Figure 3.9 Turbulence and near Wall Function .............................................................................................. 37
Figure 3.10 Schematics of hybrid LES-RANS scheme (upper) and two-layer zonal scheme
(lower) ................................................................................................................................................................................. 38
Figure 3.11 Boundary Layer Development on High Pressure Turbine Blades (Mayle)................. 41
Figure 3.12 Transition on a Low‐Pressure Turbine Airfoil at Various Reynolds Number
(Courtesy of lieva) ........................................................................................................................................................... 42
Figure 3.13 Static Pressure Distribution (Courtesy of lieva) .................................................................... 43
Figure 4.1 Schematic Representation of the Computational Domain ................................................... 49
Figure 4.2 Variation of lift Coefficient with Time ........................................................................................... 49
Figure 4.3 Comparison of Wall Pressure Coefficient (Cp) .......................................................................... 50
5

Figure 4.4 Streamline Patterns at Four Different Time Intervals for one Vortex Shedding
Cycle ...................................................................................................................................................................................... 50
Figure 4.5 Difference Between the Filtered Velocity and the Instantaneous Velocity .................. 51
Figure 4.6 Vorticity Prompted by the Wake Passing Cycle ........................................................................ 54
Figure 4.7 Time Averaged (RANS) vs Instantaneous (DES) Simulation Over a Backup Step ...... 55
Figure 4.8 Integrated RANS-LES Computations in Gas Turbines: Compressor-Diffuser, ............. 57
Figure 4.9 Compressor and combustor: RANS and LES axial velocity, mid-passage (Courtesy of
Medic et al.) ........................................................................................................................................................................ 58
Figure 4.10 DES and Effect of Grid Density on the Wake Flow ............................................................... 59
Figure 4.11 Vorticity contours from spectral DNS at two Reynolds numbers (Spalart et al.
2008). ................................................................................................................................................................................... 61
Figure 4.12 Sound Radiated by a Mach 1.9 Circular Jet............................................................................... 62
Figure 4.13 Instantaneous Contours of Stream-Wise Density Gradients from DNS ....................... 63
Figure 4.14 Stream Lines in Square Channel with Nonlinear Constitutive –( Courtesy of
[speziale, 1987]) .............................................................................................................................................................. 63
Figure 4.15 Simulation of Flow Past Circular Cylinder by Various Approaches (Shur et al.,
1996; Travin et al., 2000) ............................................................................................................................................. 64
6

Preface
This note is intended for all undergraduate, graduate, and scholars of Turbulence. It is not completed
and never claims to be as such. Therefore, all the comments are greatly appreciated. In assembling
this, I was influenced with sources from my textbooks, papers, and materials that I deemed to be
important. At best, it could be used as a reference. I also would like to express my appreciation to
several people who have given thoughts and time to the development of this article. Special thanks
should be forwarded to the authors whose papers seemed relevant to topics, and consequently, it
appears here©. Finally I would like to thank my wife, Sudabeh for her understanding and the hours
she relinquished to me. Their continuous support and encouragement are greatly appreciated.

Ideen Sadrehaghighi
June 2018
7

1 Introduction
1.1 Background
Turbulence has been the victim of many colorful descriptions over the years, from Lamb’s (1916)
scholarly “chief outstanding difficulty of our subject” to Bradshaw’s (1994) inspired “invention of the
Devil on the seventh day of creation.” This apparent frustration results largely from the mixture of
chaos and order and the wide range of length and time scales that turbulent flows possess1. The three
key elements of CFD are algorithm development, grid generation and turbulence modelling.
Turbulence is inherently three-dimensional and time dependent, and an enormous amount of
information is thus required to completely describe a Introduction Turbulent Flow. This is beyond the
capability of the existing computers for virtually all practical flows. Thus, some kind of approximate
and statistical method, called a turbulence model, is needed.
Complexity of different turbulence models may vary strongly depends on the details one wants to
observe and investigate by carrying out such numerical simulations. N-S equation is inherently
nonlinear, time-dependent, three-dimensional PDE. Turbulence could be thought of as instability of
laminar flow that occurs at high Reynolds numbers. Such instabilities origin form interactions
between non-linear inertial terms and viscous terms in N-S equation. These interactions are
rotational, fully time-dependent and fully three-dimensional. Rotational and three-dimensional
interactions are mutually connected via vortex stretching. Vortex stretching is not possible in two
dimensional space. That is also why no satisfactory two-dimensional approximations for turbulent
phenomena are available. Furthermore turbulence is thought of as random process in time.
Therefore no deterministic approach is possible. Certain properties could be learned about
turbulence using statistical methods. These introduce certain correlation functions among flow
variables. However it is impossible to determine these correlations in advance2.
Another important feature of a turbulent flow is that vortex structures move along the flow. Their
lifetime is usually very long. Hence certain turbulent quantities cannot be specified as local. This
simply means that upstream history of the flow is also important of great importance. In short, the
turbulence is severely restriction the calculation of CFD in aerospace design process where the
inability to reliably predict turbulent flows with significant regions of separation3.
Presently, turbulence modelling based on Reynolds-Averaged Navier Stokes (RANS) equations is the
most common and practical approach for turbulence simulation. RANS are time-averaged
modification of Navier-Stokes equations and turbulence models are semi-empirical mathematical
relations that are used to predict the general effect of turbulence. The objective of turbulence
modelling is to develop equations that will predict the time-averaged velocity, pressure, and
temperature fields without calculating the complete turbulent flow pattern as a function of time.
Unfortunately, there is no single universally accepted turbulence model that works for all flows and
all regimes. Therefore, users have to use engineering judgement to choose from a number of different
alternatives since the accuracy and effectiveness of each model varies depending on the application.
Turbulence” was already recognized as a distinct fluid behavior more than 500 years ago4. Figure
1.1 shows a sketch of L. Da Vinci, related to observations of free‐stream turbulence. Turbulence can
be discovered in our everyday life and surrounding phenomena such as ocean waves, wind storms

1 P., Moin and K., Mahesh, “Direct Numerical Simulation: A Tool in Turbulence Research”, Annual Rev. Fluid Mech.
1998. 30:539–78.
2 J., SODJA, “Turbulence models in CFD”, University of Ljubljana, Faculty for mathematics and physics,

Department of physics, March 2007.


3 Dimitri Mavriplis, “Exa-scale Opportunities for Aerospace Engineering“, Department of Mechanical Engineering

University of Wyoming and the Vision CFD2030 Team.


4 McDonough JM. Introductory Lecturers on Turbulence, Physics, Mathematics and Modeling. Lexington, KY:

Departments of Mechanical Engineering and Mathematics University of Kentucky. 2004/2007.


8

and smoke coming up from a


chimney, among many others.
Most flows observed in nature
and existing in various machines
and aggregates, such as pumps,
fans, turbines, dryers, cyclones,
stirred vessels, propulsion
systems and their elements
(propellers as an example),
combustion chambers, and
many others, are turbulent too.
In particular, a turbulent flow
can be expected to exhibit all of
the following specifics:
disorganized, chaotic, seemingly
random behavior; non- Figure 1.1 A sketch of turbulence by Leonardo Da Vinci,
repeatability in its structures; picture taken from (Bingham)
various and large range of length
and timescales; enhanced
mixing and dissipation, depending on the viscosity; three dimensionality, time dependence and
rotationally; intermittency in both space and time5.

1.2 Physical Description


When a sub volume of fluid is characterized by excessive kinetic energy, which is higher than the
dampening effect of the fluid’s viscosity, in other words, when very high Re number is realized in that
sub-volume, turbulence will appear. That is why turbulence is easier for observation in low viscosity
fluids, but more difficult for observation in highly viscous fluids. If Reynolds is less than a definite
critical value, damping friction forces prevent turbulent movement and the flow is laminar. In a
turbulent flow, vortex structures of various sizes and frequencies could be found. Large vortex
structures, influenced by the domain boundaries and the global flow field, break up into smaller
structures, characterized of higher frequencies. Small vortex structures are characterized by less
frequency. However, due to the flow aerodynamic character and domain boundaries, small vortices
can form bigger vortex structures and vice versa. At the same time, vortices in a volume, being at
continuous interaction, can exchange energy among them, change their energy levels and travel in
the flow volume, changing their so‐called “mixing length.”
Vortices, which contain the highest amount of kinetic energy, are described by the Taylor scale. In
the inertial range, the vortex breakup can be described by inertial effects, thus viscous effects are
negligible. Small vortices contain a low amount of energy but contribute mostly to the dissipation.
The smallest turbulent vortices are defined by the Kolmogorov microscale. In general terms, in
turbulent flow, unsteady vortices appear of many sizes which interact with each other, exchanging
energy, as a result drag increases due to friction effects. The level of turbulence has significant impact
on the stability of boundary and shear layers. High stream turbulence scales can contribute to earlier
laminar‐turbulent transition. Small turbulence scales, observed in the boundary layer, are very
important for the skin friction levels and can exert the separation due to adverse pressure gradients.
Thus, separated boundary layer and vortices lead to turbulent fluctuations and increase the level of
unsteadiness. Shear layers, between the separated and main flow, act as cores of further developed
turbulence levels and turbulent scales. Turbulence of flow through turbine channels is a prerequisite

5McDonough JM. “Introductory Lecturers on Turbulence, Physics, Mathematics and Modeling”. Lexington, KY:
Departments of Mechanical Engineering and Mathematics University of Kentucky, 2004/2007.
9

to problems in blades’ streaming; they lead to less levels of aerodynamic efficiency; changes in flow
regimes; significant pressure fluctuations, causing vibrations; and variable forces acting on blade
surfaces, among many others. Aforementioned, lead to worse efficiency, possible blades destruction,
and problems related to other processes and elements, which make part of the turbine aggregate and
installation.
10

2 Turbulence Essentials
2.1 Physical Perspectives
While the fluid elements are smooth and regular for Laminar flow pass a circle as shown in Figure
2.1 (a), the Turbulent flow are irregular in which various quantities show a random variation with
time and space as in Figure 2.1 (b). Therefore, a statically distinct averaging of values can be
distinguished. Also, because of this agitated motion in turbulent flow, the higher energy fluid
elements from the outer region of the flow are pumped close to surface. The diffusion rate of a scalar
quantity is usually greater in a turbulent flow than in a laminar. As the result, the frictional effects
are more severe for a turbulent flow. An outstanding feature of turbulent flow, as opposite to laminar
flow, is that molecules move in chaotic fashion along complex irregular path. The strong chaotic motion
causes the various layers of fluid to mix together intensively. Because of increase momentum and
energy exchange between the molecules and solid walls, turbulent flow leads at some conditions to
higher skin friction and heat transfer as compared to laminar case. It could be argued as that because
of the agitated motion in turbulent flow, the higher energy fluid elements from the outer regions of
flow are pumped close to surface, and hence, average flow velocity near solid surface is larger for
turbulent flow in comparison to laminar as depicted in following pages as Figure 2.3.

(a) Laminar flow (b) Turbulent flow


Figure 2.1 Flow Pattern Pass a Circle

2.2 Components Attributing to Complexity of Physics in Turbulence


Before plunging into the mathematics of turbulence, it is worthwhile to first discuss physical aspects
of the phenomenon. It is a rather interesting discussion which encompasses most of the physical flow.
It is intended for interest parties with full text is available in book by6. The following discussion is not
intended as a complete description of this complex topic. Rather, we focus upon a few features of
interest in engineering applications, and in construction of a mathematical model. In 1937, Taylor
and von Karman proposed the following definition of turbulence: "Turbulence is an irregular motion
which in general makes its appearance in fluids, gaseous or liquid, when they flow past solid surfaces
(boundary conditions) or even when neighboring streams of the same fluid flow past or over one another
(i,e. Jet flow)." It is characterized by the presence of a large range of excited length and time scales.

6 David C. Wilcox, ”Turbulence Modeling for CFD”, 1993, 1994 by DCW Industries, Inc.
11

The irregular nature of turbulence stands in contrast to laminar motion, because the fluid was
imagined to flow in smooth laminae, or layers. Virtually all flows of practical engineering interest are
turbulent. Turbulent flows always occur when the Reynolds number is large. For slightly viscous
fluids such as water and air, large Reynolds number corresponds to anything stronger than a small
swirl or a puff of wind. Careful analysis of solutions to the Navier-Stokes equation, or more typically
to its boundary-layer form, show that turbulence develops as an instability of laminar flow. The
features contributing to complexity of turbulence flow are;

 Enhance Diffusion
 Stability Criteria
 Time Dependency
 Eddies and Spectral Length
 Non-Linearity effects
 Separation

Enhanced Diffusion
Perhaps the most important feature of turbulence from an engineering point of view is its enhanced
diffusivity. Turbulent diffusion greatly enhances the transfer of mass, momentum and energy.
Apparent stresses often develop in turbulent flows that are several orders of magnitude larger than
in corresponding laminar flows.

Stability Criteria
To analyze the stability of laminar flows, virtually all methods begin by linearizing the equations of
motion. Although some degree of success can be achieved in predicting the onset of instabilities that
ultimately lead to turbulence with linear theories, the inherent nonlinearity of the Navier-Stokes
equation precludes a complete analytical description of the actual transition process, let alone the
fully-turbulent state. For a real (i.e., viscous) fluid, the instabilities result from interaction between
the Navier-Stokes equation's nonlinear inertial terms and viscous terms. The interaction is very
complex because it is rotational, fully 3-D and time dependent. The strongly rotational nature of
turbulence goes hand-in-hand with its three dimensionality. Vigorous stretching of vortex lines is
required to maintain the ever-present fluctuating vorticity in a turbulent flow. Vortex stretching is
absent in two-dimensional flows so that turbulence must be three dimensional. This inherent 3-D
means there are no satisfactory 2-D approximations and this is one of the reasons turbulence remains
the most noteworthy unsolved scientific problem of the twentieth century.

Time Dependency
The time-dependent nature of turbulence also contributes to its intractability. The additional
complexity goes beyond the introduction of an additional dimension. Turbulence is characterized by
random fluctuations thus obviating a deterministic approach to the problem. Rather, we must use
statistical methods. On the one hand, this aspect is not really a problem from the engineer's view.
Even if we had a complete time history of a turbulent flow, we would usually integrate the flow
properties of interest over time to extract time-averages. On the other hand, time averaging
operations lead to statistical correlations in the equations of motion that cannot be determined a
priori. This is the classical closure problem, which is the primary focus of this text. In principle, the
time-dependent, three-dimensional Navier-Stokes equation contains all of the physics of a given
turbulent flow. That this is true follows from the fact that turbulence is a continuum phenomenon. As
noted, "Even the smallest scales occurring in a turbulent flow are ordinarily far larger than any
molecular length scale”. Nevertheless, the smallest scales of turbulence are still extremely small.
They are generally many orders of magnitude smaller than the largest scales of turbulence, the latter
12

being of the same order of magnitude as the dimension of the object about which the fluid is flowing.
Furthermore, the ratio of smallest to largest scales decreases rapidly as the Reynolds number
increases. To make an accurate numerical simulation (i.e., a full time dependent 3D solution) of a
turbulent flow, all physically relevant scales must be resolved. While more and more progress is
being made with such simulations, computers of the early 1990's have insufficient memory and speed
to solve any turbulent flow problem of practical interest.

Eddies and Spectral Length


Turbulence consists of a continuous spectrum of scales ranging from largest to smallest, as opposed
to a discrete set of scales. In order to visualize a turbulent flow with a spectrum of scales we often
refer to turbulent eddies. A turbulent eddy can be thought of as a local swirling motion whose
characteristic dimension is the local
turbulence scale. Figure 2.2 shows as the
flow above the boundary layer has a steady
velocity U; the eddies move at randomly
fluctuating velocities of the order of a tenth
of U. The largest eddy size (l) is comparable
to the boundary-layer thickness (δ). The
interface and the flow above the boundary
is quite sharp [Corrsin and Kistler (1954)].
Eddies overlap in space, large ones carrying
smaller ones. Turbulence features a
cascading process whereby, as the Figure 2.2 Large Eddies in a Turbulent Boundary Layer
turbulence decays, its kinetic energy
transfers from larger eddies to smaller eddy. Ultimately, the smallest eddies dissipate into heat
through the action of molecular viscosity. Thus, we observe that turbulent flows are always
dissipative.
An especially striking feature of a turbulent shear flow is the way large bodies of fluid migrate across
the flow, carrying smaller-scale disturbances with them. The arrival of these large eddies near the
interface between the turbulent region and non-turbulent fluid distorts the interface into a highly
convoluted shape. In addition to migrating across the flow, they have a lifetime so long that they
persist for distances as much as 30 times the width of the flow [Bradshaw (1972)]. Hence, the
turbulent stresses at a given position depend upon upstream history and cannot be uniquely
specified in terms of the local strain-rate tensor as in laminar flow.

Non-Linearity Effects
The nonlinearity of the Navier-Stokes equation leads to interactions between fluctuations of differing
wavelengths and directions. As discussed above, the wavelengths of the motion usually extend all the
way from a maximum comparable to the width of the flow to a minimum fixed by viscous dissipation
of energy. The main physical process that spreads the motion over a wide range of wavelengths is
vortex stretching. The turbulence gains energy if the vortex elements are primarily oriented in a
direction in which the mean velocity gradients can stretch them. Most importantly, wavelengths that
are not too small compared to the mean-flow width interact most strongly with the mean flow.
Consequently, the larger-scale turbulent motion carries most of the energy and is mainly responsible
for the enhanced diffusivity and attending stresses. In turn, the larger eddies randomly stretch the
vortex elements that comprise the smaller eddies, cascading energy to them7.

7 David C. Wilcox, “Turbulence Modeling for CFD”, Copyright © 1993, 1994 by DCW Industries.
13

Separation and Drag Reduction


Although the frictional effects are more severe in
turbulent flow, both shear stress and aerodynamic
heating are larger for the turbulent flow in comparison
to laminar8. However, turbulent flow has a major
redeeming value; because the energy of the fluid
elements close to surface is larger, it does not
separates from the surface as readily as a laminar flow.
Simply, if a flow is turbulent, it is less likely to separate
from the body surface. And if flow separates, the
separation region is smaller than those for laminar. As
the result the pressure drag Dp will be smaller in
turbulent flow. This leads to great compromise in
aerodynamics. In general, if the body is slender, the
friction drag, Df is much greater than Dp. For this case
since Df is smaller in laminar flow (lower velocity →
lower viscous drag; see Figure 2.3), then laminar flow
is desirable for slender bodies (airfoils). In contrast, for
a blunt body, Dp is much greater than Df. For this case, Figure 2.3 Velocity Profile for Laminar vs
Turbulent Flows
since Dp is smaller in turbulent flow, then turbulent
flow is desirable (gulf balls).

2.3 Remarks Regarding Transition Phenomenon


The transition from laminar to turbulent is not subtle and consists of several processes, each subject
to intense research interests on their own right. The transition and stability of the laminar flow seem
to be dependent on a critical value of Reynolds number (Error! Reference source not found.) as

ρ  V x cr
Re cr   2100 Eq. 2.1
μ

Figure 2.4 Turbulence Transition Region

8 Anderson, John D. 1984: “Fundamentals of Aerodynamics”, McGraw Hills Inc.


14

The equations governing a turbulent flow are precisely the same as for laminar flow; however,
solutions is clearly more complicated in this region. That is due to the introduction of new terms and
issues with closure. Two general approach could be envisioned. First a more direct approach with
extreme spatial discretization to capture all the flow eddies near the wall. This is prohibitly CPU
intensive, even with current computing powers. To counter that, the use of some modeling (empirical
or otherwise) in the vicinity of wall region is advocated. Rest of this report organized as subsequent.
Since subject of Turbulence is very involved, therefore, the physical aspects of formulation, namely
Reynolds Stress formulation, and how to derive them represented first. Followed by different aspects
of Turbulence modeling.
The numerous factors contributing to transition from laminar to turbulent flow in a fluid. The intent
has been to provide general background information on the various transition phenomena rather
than to make a study of the problem in depth. Included are the effects on transition of such factors as
pressure gradient, surface temperature, Mach number, and 2 and 3D types of surface roughness
[A. L, Braslow]9. The effect of wall heat transfer was investigated by [Shadloo and Hadjadj]10 through
high-resolution DNS , among other factors. As in the case of adiabatic wall (Error! Reference source
not found.-a), stream-wise-elongated streaks are visible in the laminar and transitional regions for
both heated (Error! Reference source not found.-b) and cooled cases (Error! Reference source not
found.-c). However, they locally break down and create turbulent spots further downstream for the
heated case when compared with the adiabatic wall. Therefore, the wall heating stabilizes the flow
and postpones the transition.

(A) Adiadatic Wall


(No Heating)

(B) Tw/Tr = 1.5


(Heated)

(C) Tw/Tr = 0.75


(Cooled)

Figure 2.5 Snapshots of the Stream-Wise Velocity Component for (A) Adiabatic Wall , (B) Heated
Wall, and (C) Cooled Wall

9 Albert L, Braslow, “A Review Of Factors Affecting Boundary-Layer Transition”, NASA TN D-3384.


10 M. S. Shadloo & A. Hadjadj, “Laminar-turbulent transition in supersonic boundary layers with surface heat
transfer: A numerical study”, Numerical Heat Transfer, 2017.
15

The transition phenomenon, to model and resolve, was also described extensively by [lieva]11.
Prediction of the onset of boundary layer transition is one of the most important concerns in the area
of fluid mechanics. There is a great interest to transition as it plays a major role in many engineering
applications and raises important questions to the flow physics, also could serve as an ingesting
example for determinism and chaos. The so‐called viscous instability of a laminar boundary layer
was for the first time taken into account and studied by Tollmien. Under low free‐stream turbulence
conditions, instability is observed in the case of two‐dimensional unstable Tollmien‐Schlichting
waves are formed and propagate in the stream wise direction. These waves lead to additional 3D
aerodynamic effects to appear in the flow structure, such as peaks, stronger secondary flow effects,
hairpin vortices and transition effects. Turbulent spots are formed in the regions of vorticity peaks
and can develop to continuously spreading turbulence. A turbulent spot model to describe the
specifics of a transitional flow is proposed in . Later, turbulent spots generated over a flat plate
surfaces, without imposed pressure gradients, were also visualized .
Recently, scientists have been working on more accurate transition length predictions, based on
measurement of transition length in a field of adverse pressure gradients and of triggered turbulent
spots. It was found that spot characteristics, in the case of adverse pressure gradients, are different
from those formed in the case of zero or favorable pressure gradients. Also, it became clear that in
the presence of adverse pressure gradient, a spot can be formed at the center of a highly amplified
transverse waves and is convected at lower velocity than under a zero pressure gradient. Laminar to
turbulent transition is proved as a phenomenon, which seriously affects the efficiency performance
of various machines. The transition effects contribute to additional drag and lift forces, also heat
fluxes that are crucial for overall working principles of different types of machines and installations.

Transition Types
Speaking about transition, one must stress that there are different types of transition. There are:

2.3.1.1 Natural Transition


The first one is called “natural” transition. It begins with a weak instability in the laminar boundary
layer, as was described years ago by [Tollmien & Schlichting] , next proceeds through various stages
of amplified instability to fully turbulent flow.

2.3.1.2 Bypass Transition


The second type of transition is the “bypass”” transition, defined by [Morkovin] . Bypass transition
is caused by high levels of disturbances in the external flow (such as free‐stream turbulence) and can
completely bypass “natural transition.” As it is related to the “bypass” transition, some researchers
presumed that it could appear as an instantaneous turbulent breakdown with zero length of
transitional flow. However, the bypass transition does not always exclude instability processes. Only
the long region of two dimensional wave amplification preceding the appearance of three‐
dimensional disturbances (span wise periodicity) in low turbulence flow is bypassed. In more detail,
during the phase of the so‐called “bypass transition,” not all specific laminar breakdown processes
would be recognized. Separated‐flow transition occurs when a laminar boundary layer separates and
transitions in the free shear layer, above the already formed bubble. Transition due to separated flow
could develop at the leading edge and close to the place where minimum pressure on the suction
surface sides is formed. This type of transition is extremely harmful for low pressure turbines and
leads to early separation of the boundary layer.

11 Galina Ilieva Ilieva, “On Turbulence and its


Effects on Aerodynamics of Flow through Turbine Stages”, Technical
University, Varna, Bulgaria, Intech Open 2017. http://dx.doi.org/10.5772/intechopen.68205
16

2.3.1.3 Reverse Transition


There is another type of transition, reverse transition, which is known as “relaminarization.”
Relaminarization is possible to appear at places where a previously turbulent or transitional region
is affected by strong favorable pressure gradients, and as a result of that, it transfers to laminar again.
Depending on the profile section geometry and the flow regime, near the leading edge, laminar flow
followed by a wake‐induced or shock‐induced transition could be visualized. Last described
phenomenon could be replaced by a relaminarization with subsequent transition to turbulence,
occurring at multiple locations simultaneously.

2.3.1.4 Separated Flow Transition


The third type of transition is the so called “separated‐flow” transition. It exists in a separated laminar
boundary layer and may or may not reveals some instabilities of the Tollmien‐Schlichting type. The
phenomenon of “separated‐flow transition” could arise after the so‐called boundary layer trip wires
as a result of laminar separation under strong adverse pressure gradients. Thus, the flow can reattach
as turbulent, forming laminar separation/turbulent‐reattachment “bubble”, on the surface under
consideration. In gas turbine stages, the transition of separated flow could be seen in the so‐called
over speed region close to the leading edge of the profile, over the convex or concave side, or both,
and near the place where minimum pressure on the convex side is observed. What will be the bubble
size depends on the transition process within the free shear layer and may involve all of the stages
for a natural transition type. For bubbles with bigger size and characterized by low free‐stream
turbulence levels, flow in the bubble is dominantly laminar and instabilities could be observed. Big
bubbles, along the blade surfaces, produce losses and act as a prerequisite to exit flow angles
deviation. Small bubble configurations are an effective way to increase the turbulence levels and can
possibly control the blade aerodynamics.

2.3.1.5 Effect of Geometry


The effect of geometry curvature and/or streamlines curvature on the transition was studied. In , it
was found that a laminar boundary layer formed over a concave surface becomes unstable as a result
of acting centrifugal forces, three‐dimensional disturbances and stream wise vortices presented in
the boundary layer. The increase in transition Reynolds number is caused by the Görtler vortices,
which can increase velocity gradients near the wall and thus can delay the transition. For highly
curved surfaces, this effect dominates that one caused by turbulence. A concave curvature can
decrease or can increase the transition The Reynolds number depends on the turbulence intensity
and the curvature. Heating or cooling can seriously affect boundary layer transition at low free‐
stream turbulence. Heat transfer through a laminar boundary layer formed over the concave blade
surface is influenced both by Taylor‐Görtler vortices and the main flow turbulence levels. Transition
occurs when these factors surpass the tendency of boundary layer to remain laminar in the presence
of higher pressure gradients. If spot production is not affected by the heat transfer, at high free‐
stream turbulence intensities, transition would not be observed.

2.3.1.6 Boundary Heat Transfer


Heating or cooling can seriously affect boundary layer transition at low free‐stream turbulence. Heat
transfer through a laminar boundary layer formed over the concave blade surface is influenced both
by Taylor‐Görtler vortices and the main flow turbulence levels. Transition occurs when these factors
surpass the tendency of boundary layer to remain laminar in the presence of higher pressure
gradients. If spot production is not affected by the heat transfer, at high free‐stream turbulence
intensities, transition would not be observed. Film cooling affects the boundary layer formed on the
streamed surfaces. At places where cooling fluid is injected, holes are usually much larger than the
boundary layer thickness, thus the injection of coolant through these holes disrupts the flow close to
the surfaces and provides higher turbulence levels within the downstream developing boundary
17

layer [63]. Therefore, it may be said that film cooling effect is to “trip” a laminar boundary layer and
initiates transition to turbulence. In the case of acceleration, sufficient to cause reverse transition in
the downstream direction of the injection, the heat transfer intensity approaches that for laminar
flow. This implies that even though injection can initiate transition, a subsequent strong acceleration
can cause the flow to become laminar again. Such a situation is common for film‐cooled blades of first
gas turbine stages. Heat transfer measurements on a stator vane, presented in [85, 86], indicated that
the behavior of the boundary layer transition along the suction side of the vane showed dependency
to the film‐cooling injection place.

What Triggers Transition?


In general, the turbulence Reynolds number increases with increase in acceleration or with decrease
in the free‐stream turbulence levels. The effect of acceleration is significant for low turbulence levels.
However, for turbulence levels found in gas turbine stages, it has a negligible value. In the case of high
turbulence levels, the transition onset is controlled by the free‐stream turbulence. For accelerating
flows, length of transition is variable for thermal and momentum boundary layers. The local point
where boundary layer can separate is not in a close relation to the levels of free‐stream turbulence.
In the case of observed increase in levels of free‐stream turbulence, the separation bubble size
decreases and turbulent transition point moves in upstream in regard to the flow direction. The
separation bubble decreases with increase of the Reynolds number .
At high free‐stream turbulence intensity, the streak structures can be observed in the upstream
direction and are related to the place of boundary layer separation, showing that bypass transition
of an attached boundary layer can be realized at the high Reynolds number. In the case of low free‐
stream turbulence intensity, velocity fluctuations are seen within the shear layer of the separation
bubble. The roughness over streamed surfaces is non uniform and characterized by significant
variations in stream wise and span wise directions. Modifications in the behavior of the boundary
layer due to the presence of definite values of surface roughness can decrease the aerodynamic
efficiency and can also increase the heat transfer). Levels of heat transfer may also be affected by
changes in material properties or in the case of eroded or broken protective coatings.
18

3 Reynolds Averaging and RANS Modeling (Linear Modeling)


3.1 Time Averaging
Since Most of engineers does not concern with instantaneous values but rather deal with averages.
Therefore, the Reynolds Averaged N-S equations (RANS) are derived by decomposing the dependent
variables in conservative equations into time mean and fluctuating components. Two types of
averaging are greatly used, the classical Time (Reynolds) average, and the Mass Weighted average. For
flows in which density fluctuating could be neglected, the two formulations becomes identical. The
time averaging procedure defined by:

t 0  Δt
1
f
Δt  f dt
t0
and f   0 , u i  u i  ui , ρ  ρ  ρ , p  p  p , T  T  T Eq. 3.1

It is required that Δt be large compared to the period of random fluctuations associated with
turbulence, but small with respect to the time constant for any slow variations in flow field associated
with ordinary unsteady flows. In conventional Reynolds decomposition, the randomly changing flow
variations are replaced by the time average plus fluctuations (see Figure 3.1) and would be discuss
later for Unsteady RANS (URANS) formulations. The resulting of decomposition of variables and time
averaging the entire equation, the continuity, momentum, and energy yield to

ρ 
M ass :  (ρ u j  ρu j )
t x j
 
M omentum : (ρ u i  ρu i )  (ρ u i u j  u i ρu j ) 
t x j
p 
  ( τ  u ρu   ρ u  u   ρu  u  )
x x ij j i i j i j
i j
 (c ρT  c ρ T)
p p   
Energy :   ρc u T  
t x  p j 
j
p p p    T 
 uj  u j  k  ρc p Tu j  c p ρTu j   
t x j x j x j  x j 

ui u 
where   τij  τij i
x j x j
 u u j  2 u k  1 i j
τij  μ  i    δ ij  and δ ij  
 x j x i  3 x k  0 i j
Eq. 3.2
19

Figure 3.1 Relationship between averaged variables; a) Steady flow b) Un-Steady flow

Un-Steady RANS (URANS)


For unsteady analysis, the Unsteady RANS could be used where of course the unsteady term is
retained in averaging as well. In essence, there are now two time scale involved where the Reynolds
decomposition of U would be as before an average and fluctuated value as


1
U
  U( ) d
0
, U  U  u  Eq. 3.3

Note that the dependent variables are now not only a function of space but also a function of time as
well. Be advised that this concept of time, is different from the time step (t) of mean value. In essence
if you solve your equations with one global time step, which is used in every cell, and if value of the
time step is small enough then you will be able to capture fluctuations, or unsteady behavior in the
MEAN quantities. In other words your solution is time accurate. Steady RANS, or RANS, marches the
solution with a local optimized time step for each cell, and hence is not time accurate, you will get a
faster solution, and it will be steady state. For URANS we have:

Ui  Ui (x, y,z, ) and uiuj  uiuj (x, y,z, ) Eq. 3.4

Even if the results from URANS are unsteady, one is often interested only in time averaged flow as
denoted as <Ū>, which means that we can decompose the results from URANS as a time averaged
part <Ū>, a resolved fluctuation uˈ, and the modeled turbulent fluctuation, u̎ , i.e.

U  U  u   U  u  u Eq. 3.5

3.2 Reynolds Averaged Naiver-Stokes (RANS)


The Reynolds Averaged Navier-Stokes equations (RANS) are obtained from the continuity and
momentum equations by taking the time average of all the terms in the equations. The continuity
equation does not change since it is linear in terms of the velocity. However, the momentum equation
is non-linear, which means that all the fluctuating components do not vanish. An extra term, called
Reynolds stress uiuj, appears in the momentum equation. The result can be written where the
20

overbar has been dropped from the mean values. This convention will be used throughout here for
obvious reasons. Eq. 3.6 presents the fundamental problem of turbulence. In order to compute all
the mean-flow properties of the turbulent flow we need a reasonably accurate way to compute the
Reynolds stress uʹiuʹj. This is the fundamental reason for the need of the turbulence models. The
complete set of the Reynolds Averaged Navier-Stokes (RANS) equations are not presented here
because they can be found from many references12. Scalar transport equations are also needed, for
example to describe the transport of the concentration of species or the mass fraction of species.
Their exact formulation can be found in13. For incompressible flow with constant properties and no
body force the momentum and energy.

Closure
Problem

Eq. 3.6

Comments on Reynolds Averaged Naiver-Stokes (RANS)


At the first glance the Reynolds equations seemed to be quite complicated to solve in turbulent flow.
Certainly a major problem in fluid mechanic is that they are more equation can be written than
solved. Fortunately, for many important flows, the Reynolds equations can be simplified. But before
turning into the task of simplification, it is prudent to examine the equation themselves. By
considering the incompressible flow, the momentum equation can be written as:

Du̅i ∂p̅ ∂(τ̅ij )Lam ∂(τ̅ij )Turb


ρ = − + +

Dt ∂x
⏟i ⏟ ∂xj ⏟ ∂xj
Particle Acceleration Mean Pressure Laminar like Stress Apperant Stree Gradient
of Mean Motion Gradient Gradient due to Transport of Momentum
by Turbulent Fluctuation
where (τ̅ij ) ̅̅̅̅̅
= −ρu ′ u′
i j Similarly for Energy Equation:
Turb

12 P. Kaurinkoski and A. Hellsten, FINFLO: The Parallel Multi-Block Flow Solver, Report A-17, Laboratory of
Aerodynamics, Helsinki University of Technology, Espoo, Finland, 1998.
13 P. Kaurinkoski, “Development of an Equation of State for an Arbitrary Mixture of Thermally Perfect Gases to the

FINFLO flow solver”, Report No B-48, Series B, Laboratory of Aerodynamics, Helsinki University of Technology,
Finland, 1995.
21

∂ ̅
∂T ∂
−(∇. q)Lam = (k ) and − (∇. q)Turb = (ρCp ̅̅̅̅̅̅
T ′ u′j )
∂xj ∂xj ∂xj
Eq. 3.7
The Reynolds equation cannot be solved in the form given because the new apparent turbulent
stresses and heat-flux quantities must be viewed as new unknowns. To proceed further, we need to
find additional equations involving the new unknowns to make assumption regarding the
relationship between the new apparent turbulence quantities and time-mean flow variables. This is
known as closure problem which is most commonly handled through turbulence modeling to be
discussed next.

3.3 Closure Problem (Boussinesq Assumption)


A linear relation between turbulent shearing stresses and rate of mean strain suggested by
[Boussinesq] which is the fundamental closure for most eddy viscosity models been used.

 u i u j  2  u k 
 ρui uj  μ T     δij  μ T  ρ κ  κ  ui ui /2 Eq. 3.8
 x x  3  x 
 j i   k 

Where μT is the turbulent viscosity, κ is the kinematic energy of turbulence are key modeling aspects
associated with so called turbulence closure problem. Reynolds stresses with deformation rate, and
are related to viscosity, mean rate of deformation, and turbulent kinetic energy with Boussinesq’s
proposal expressed can be used to calculate Reynold’s stresses in the final step of turbulence
modelling. It is seen from this equation that the Reynold’s stresses are considered proportional to
the dissipation rate reduced by the eddy turbulent kinetic energy. (The Kronecker delta ensures that
the normal Reynolds stresses are each appropriately accounted for). It can also be seen that the
kinetic energy allocates an equal third for each normal stress component (isotropic assumption).
This is the reason for the inherent inaccuracy of the κ-ε model, making it incapable of describing
anisotropic flow. Other scalar flow properties such as mass and heat can also be modelled using time-
averaged values. To obtain mathematical closure, the Reynolds stress terms must be related to mean
flow properties either empirically or through flow model which allows calculation of this
relationship. References are also sometimes made to the order of the closure. According to this
terminology, a 1st order closure evaluates the Reynolds stresses through functions of the mean
velocity and geometry alone, as the case above (linear). A 2 nd order closure employs a solution to a
modeled form of transport partial differential equations for one or more of the characteristics of
turbulence (non-linear). More complicated quadratic or cubic closures are also available through
literature.

3.4 Turbulence Models


Turbulence models to close the Reynolds equations can be divided originally into three groups in
general.

 First models which (directly) use the Boussinesq assumption. Most models currently
employed in engineering are this type. Experimental evidence indicates this valid in many
circumstances.
 Second are models using the effect of closure to the Reynolds equation without this
assumption.
 The third category is defined as those that are not based entirely on the Reynolds equation.
An example would be Large Eddy Simulation (LES) where an attempt is made to resolve the
22

large scale turbulent equation from first principals be numerically solving the filtered set of
equations governing this large scale. Turbulence modeling is used to approximate the effects
of the sub-grid scale (SGS) turbulence.

Mixing Length
0-Eqaution
Baldwin-Lomax

1-Equation Spalart-Almars Standard


1st - order
(linear)
κ-ε Models RNG

Reynolds Realizable
Average 2-Equation
(RANS)
SST
Reynolds Stress
κ-ω Models
Standard
Turbulence 2nd - order
Algebric Stress (Wicox)
Models (non-linear)

Quadratic &
LES DES
Cubic κ-ε

DNS

Figure 3.3 Hierarchy of Turbulence Models in General

Although such a calculation


have shown great promises, it is
computationally prohibited to
be considered as engineer in
tool14. There are other
references to turbulent
modeling such as order of
closure. According to this
terminology, a first order
closure evaluates the Reynolds
stresses through functions of
the mean velocity and geometry
alone. A second order closure
employs a solution to a modeled
from transport partial
differential equation for one or
more of the characteristics of Figure 3.2 Hierchey of Turbulence Models based on Relative
turbulence. Since this Importance of Numerics vs. Computational Costs - (Courtsy of Tenzor)

14Anderson, Dale A; Tannehill, John C; Plecher Richard H; 1984,”Computational Fluid Mechanics and Heat
Transfer”, Hemisphere Publishing Corporation.
23

straighter forward, we try to adapt this category as depicted in Figure 3.3. These are some of models
mentioned here and by no means is it exclusive. An ideal model should be with minimum amount of
complexity while capturing the essence of the relevant physics. Figure 3.2 shows the relative
importance of numeric vs. computational cost for various models using a pyramid.

Turbulent Viscosity, Velocity, and Length Scale


Before we start the derivation for κ-ε and other models, there are some common ground that we
should cover. As viscosity is the central concept for modelling the stresses, it should be noted that the
turbulent viscosity μT is expressed as:

μ T  ρu T l where uT 
2
3
 
1
3

ui ui  Eq. 3.9

It can be seen that the turbulent viscosity is a product of density and two new variables representing
turbulent velocity and turbulent length scale. Turbulent velocity uT can be described as the typical
velocity occurring in the largest eddies and can also be related to the same eddies. Turbulent kinetic
energy according to turbulent length scale is the average length. The new variables, uT and l form the
basis for the “two-equation‟ k-ε turbulence model, meaning that in addition to the RANS equations,
two more equations are required to solve for turbulent velocity and turbulent length using the model.
The length scale l (for large eddies) is used in the k-ε model to define the length scale ε (for small
eddies), for which a transport equation is used in the model, and represents the dissipation of the
turbulent kinetic energy. The dissipation is expressed as:

κ 2/3 μ T  ui  u j
ε as related to viscosity ε   Eq. 3.10
l ρ x κ x κ

As seen in the second expression for ε, dissipation of the turbulent kinetic energy κ is proportional to
the rate of deformation of eddies. Other scalar flow properties such as mass and heat can also be
modelled using time-averaged values. Similar to the turbulent momentum transport’s
proportionality to average velocity gradients, turbulent scalar transport is proportional to mean
scalar value gradients and can be expressed as


 ρu i   Γ T (9.16) Eq. 3.11
x i

Where ГT refers to turbulent (eddy) diffusivity. As can be seen from above expression, the turbulent
scalar property transport occurs with the same mechanism as in transport of momentum (mixing of
eddies, represented by ГT). For this reason, it can be assumed by Reynold’s analogy that the value of
ГT is similar to μT, the turbulent viscosity. The ratio of μT to ГT is defined as the Prandtl/Schmidt
number σT, and has a value which is normally constant with a value around unity. The next section
presents the κ-ε model and the two extra k and ε transport equations (PDEs) for closing the system
of time-averaged RANS equations. The model is based on the mechanisms causing changes to
turbulent kinetic energy (i.e. turbulent viscosity and velocity fluctuations).

3.4.1.1 Eddy Viscosity (RANS) Models


Van Driest (1956) devised a viscous damping correction for the mixing-length model. This correction
is still in use in most modern turbulence models. [Cebeci & Smith] refined the eddy viscosity/mixing-
24

length concept for better use with attached boundary layers15.

3.4.1.2 Algebraic Model (Zero Equation) - Mixing Length


This is invariably the successful this type model which uses the Boussinesq assumption. It was
originally suggested by Prandlt in 1950s a

u
μ T  ρu T l or μ T  ρl 2 Eq. 3.12
y

Where l a mixing length can be thought of as a transverse distance over which the particles maintain
their original momentum. For 3D thin shear layers, Prandtl’s equation is usually interpreted as:

1/2

2  u 
2
 w  
2

μ T  ρl       Eq. 3.13


 y   y  

This formula treats the turbulent viscosity as a scalar and gives qualitively correct results, specially
near the wall. The evaluation of l in the mixing length model varies with the type of flow being
considered, wall boundary layer, jet, wakes, etc. For flow along a solid surface (internal or external),
good results are observed by evaluating l according to

 y  /A   y( τ w /ρ w )1/2
linner  κy(1 e ) , louter  C1δ and y  Eq. 3.14
νw

Where linner predicts the inner region close to wall and Iouter exceeds linner. The κ is the von Karman
constant as 0.41 and A+ is the damping constant usually set to 26. The expression for linner is
reasonable for producing the inner law-of-the-wall region of turbulent flow and louter produces the
outer “wake-like” region. These two zones are depicted in Figure 3.8 which shows a typical velocity
distribution for an incompressible turbulent boundary layer on a smooth impermeable plate using
“law-of-the-wall” coordinates.

3.4.1.3 Baldwin-Lomax
[Baldwin & Lomax -1978] proposed an alternative algebraic model to eliminate some of the difficulty
in defining a turbulence length scale from the shear-layer thickness. It is a two-layer algebraic 0-
equation model which gives the eddy viscosity μt as a function of the local boundary layer velocity
profile. The model is suitable for high-speed flows with thin attached boundary-layers, typically
present in aerospace and turbo machinery applications. It is also commonly used in quick design
iterations where robustness is more important than capturing all details of the flow physics. The
Baldwin-Lomax model is not suitable for cases with large separated regions and significant
curvature/rotation effects.

One-Equation Model
While employing a much simpler approach than two-equation or second-order closure models, one-
equation models have been somewhat unpopular and have not showed a great deal of success. One
notable exception was the model formulated by [Bradshaw, Ferris, and Atwell -1967], whose model
was tested against the best experimental data of the day at the 1968 Stanford Conference on

15 David C. Wilcox, “Turbulence Modeling for CFD”, 1993, 1994 by DCW Industries, Inc.
25

Computation and Turbulent Boundary Layers. There has been some renewed interest in the last
several years due to the ease with which one-equation models can be solved numerically, relative to
more complex two-equation or second-order closure models. An obvious shortcoming of algebraic
methods is that μT and uT is zero at the center of for example pipe line cases. The mixing-length model
can be fixed up to overcome this using

μ T  C k ρl (κ)1/2
For 2D incompressible thin - shear layer Eq. 3.15
Dκ   μ T  κ 
2
 u  C ρ(κ) 3/2
   μ    μ T    D
Dt y  Prκ  y   y  l

Where Prκ is defined as Prandtl number for turbulence kinetic energy (≃1.0) and CD ≃ 0.164.

3.4.2.1 k - Model
The complete derivation is obtained in two equation model.

3.4.2.2 Spallart - Allmaras


The [Spalart – Allmaras] model adds a single additional variable for a Spalart - Allmaras viscosity and
does not use any wall functions; it solves the entire flow field. The model was originally developed
for aerodynamics applications and is advantageous in that it solves for only a single additional
variable. This makes it less memory-intensive than the other models that solve the flow field in the
buffer layer. Experience shows that this model does not accurately compute fields that exhibit shear
flow, separated flow, or decaying turbulence. Its advantage is that it is quite stable and shows good
convergence.

Two Equation Model


While Kolmogorov’s κ-ω model was the first two-equation model, the most extensive work has been
done by [Daly & Harlow; 1970] and [Launder & Spalding; 1972]. Launder's κ-ε model is the most
widely used two-equation turbulence model; here ε is the dissipation rate of turbulent kinetic energy.
Independently of Kolmogorov, [Saffman ; 1970] developed a κ-ω model that shows advantages to the
more well-known κ-ε model, especially for integrating through the viscous sub-layer and in flows
with adverse pressure gradients. The κ-ε model is very popular for industrial applications due to its
good convergence rate and relatively low memory requirements. It does not very accurately compute
flow fields that exhibit adverse pressure gradients, strong curvature to the flow, or jet flow. It does
perform well for external flow problems around complex geometries.
It was concluded that the algebraic and one equation models did not lend itself to application with
the complex geometries that occur with internal flow calculations. In this context, we define an
internal flow as a flow through turbomachinery blading, as opposed to the external flow that occurs
with isolated airfoil or aircraft wing flow field predictions. When dealing with internal flows, one can
overcome the limitations of a one-equation model with a two-equation model. The currently favored
two-equation models are either the 𝑘-𝜀 model or the 𝑘-𝜔 model16. Both models require one to solve
two transport equations to compute eddy viscosity as an algebraic expression of turbulent kinetic
energy (𝑘) and dissipation rate of turbulent kinetic energy (𝜀) or turbulence frequency (𝜔). In two-

16Alessandro Corsini, Giovanni Delibra, and Anthony G. Sheard, “A Critical Review of Computational Methods
and Their Application in Industrial Fan Design”, Hindawi Publishing Corporation, ISRN Mechanical Engineering,
Volume 2013, Article ID 625175.
26

equation models the first equation is typically for turbulent kinetic energy (𝑘) and the second for
either dissipation (𝜀) or turbulence frequency (𝜔). Both the 𝑘-𝜀 model and the 𝑘-𝜔 model rely on an
assumption that one can link eddy viscosity to a time and length scale that characterizes turbulence
that in turn links to the computed flow-field’s characteristics. A feature of the 𝑘-𝜀 and 𝑘-𝜔 models is
that the additional transport equations for 𝑘, 𝜀, and 𝜔 share the same form, and, therefore, for a
generic 𝜙 quantity, it reads

D   ν   
 P  ε   ν  T 
 x  Eq. 3.16
Dt x j  σ  j 

On the left-hand side is the quantity’s material derivative. On the right-hand side are one or more
production terms, a dissipation term, a diffusion term dependent on molecular viscosity, and another
given as the turbulent viscosity’s function, corrected using the Prandtl number 𝜎𝜙. The primary
difference between the 𝑘-𝜀 and 𝑘-𝜔 models is the different trend of 𝜀 and 𝜔 at the wall and the definition
of the wall boundary conditions for the same variables. When one studies normalized values of 𝜀 and
𝜔 for an attached flow, it is evident that 𝜔 is less dependent on the Reynolds number than 𝜀 in the
wall’s near vicinity.
There is a general consensus within the computational fluid dynamics community that the 𝑘-𝜀 model
better reproduces the energy cascade of large-scale structures in the main flow core, whilst the 𝑘-𝜔
model performs better near the wall17. A realization that 𝑘-𝜀 models perform better in the
main flow whilst the 𝑘-𝜔 models perform better near the wall leads to the natural conclusion that,
ideally, one would use the two models in combination. It was observed that it is possible to combine
𝑘-𝜀 and 𝑘-𝜔 models as one can reformulate every two-equation model into every other by changing
model coefficients. This realization has enabled engineers to formulate the 𝑘-𝜔 shear stress transport
(𝑘-𝜔 SST) model that solves the equation for 𝜔 near the wall and 𝜀 elsewhere. The use of two-
equation models has become established within the industrial community.

3.4.3.1 Standard Transport Equation for k


To obtain the equation for turbulent kinetic energy k, complicated algebra and rearrangements are
made to the time-averaged continuity equation and the time-averaged Navier-Stokes equations for
momentum. The mathematical manipulations are extensive, and therefore only a short description
of what is done, followed by the resulting equations is presented. The 3 continuity equations are each
multiplied by the respective fluctuating velocity component and then added together. The same
process is carried out for the Reynolds equations for momentum. The two resulting equations are
subtracted and rearranged extensively. Terms for viscous dissipation of turbulent kinetic energy and
the Boussinesq assumption related to Reynold’s stress equation, as shown below in18.

∂(ρκ) ∂(ρu̅i κ) ∂ u′i u′j ∂u̅i ̅̅̅̅̅̅̅̅̅̅


∂u′i ∂u′i
+ = ′
[ρui . ( ̅̅̅̅̅
+ p′ )] − ρu ′ u′ .
i j − μ. .
⏟∂t ⏟ ∂xi ∂x
⏟i 2 ⏟ ∂x j ⏟ j∂x ′ ∂x ′
j
1 2 3 4 5
Eq. 3.17
Terms (3) and (5) are replaced using scalar diffusion transport terms for (3) and the time-averaged
term for (5) to result in the following:

17 Alessandro Corsini, Giovanni Delibra, and Anthony G. Sheard, “A Critical Review of Computational Methods
and Their Application in Industrial Fan Design”, Hindawi Publishing Corporation, ISRN Mechanical Engineering,
Volume 2013, Article ID 625175.
18 Tennekes, H.; Lumley, J. L. “A First Course in Turbulence”, MIT Press, Cambridge, MA (1972).
27

∂(ρκ) ∂(ρu̅i κ) ∂ μT ∂κ ∂u̅i


+ = [ ̅̅̅̅̅
] − ρu ′ u′ .
i j − ρ.
⏟ ε
⏟∂t ⏟ ∂xi ∂x
⏟ i σκ ∂xi ⏟ ∂xj
5
1 2 3 4
Eq. 3.18
Where is a constant turbulent Schmidt number for κ. The terms (1)-(5) of the turbulent kinetic energy
k transport equation, can be interpreted as the following:

1. Transient term Accumulation of κ (rate of change of k).


2. Convective transport of κ by convection.
3. Diffusive transport of κ by pressure, viscous stresses, and Reynolds stresses (must be
modelled).
4. Production term Rate of production of k due from the mean flow.
5. Viscous dissipation Rate of viscous dissipation of k (must be modelled) k is a constant
(turbulent Schmidt number) of the k equation.

The above equation for transport of κ along with the equation for transport of ε, constitute the two
additional transport equations to be solved in addition to the RANS equations in the κ-ε turbulence
model. The next section presents the equation for transport of ε.

3.4.3.2 Standard Transport Equation for ε

∂(ρε) ∂ ∂ μT ∂ε ε ∂u̅i ε2
+ (ρu̅i ε) = [ ] + C1 (−ρuu. ) − C2 ρ
⏟∂t ∂xi
⏟ ∂xi σε ∂xi
⏟ κ
⏟ ∂xi ⏟ κ
1 2 3 4 5
Eq. 3.19
Similar to the transport equation for κ, the transport equation for ε includes the terms 1 - 5:

1. Accumulation, or rate of change, of ε, Cμ C1 C2 σκ σε


2. Rate of destruction of ε. 0.09 1.44 1.92 1.0 1.3
3. Diffusive transport of ε,
Table 3.1 Constant values for κ-
4. Rate of production of ε, and
ε turbulence model equations
5. Transport of ε by convection,

And C1, C2, σε are constants of the ε equation (see Table 3.1). Alternatively, we inscribe both
equations in a more compact form, using Lagrangian Derivatives terms D/Dt:

Dκ  1  μT  κ  μ T 2
   μ    Sij  ε
Dt x j  ρ  σk  x j  ρ
Dε ε μ  1  μT  ε 
 (C1 T Sij2  C 2 ε)    μ    Eq. 3.20
Dt κ ρ x j  ρ  σε  x j 
κ2 1  u u j 
where μ T  ρCμ and Sij   i  
ε 2  x j x i 
28

After obtaining κ and ε we are ready to evaluate μT as

Cμ ρ(κ) 2
μT  Eq. 3.21
ε

Despite the enthusiasm which is noted from time to time over two equation model, it is perhaps
appropriate to point out two major restriction on this type of models. Since the two equation model
basically turbulent viscosity models which assumes that the Boussinesq approximation holds.
Therefore, its validates depends to Boussinesq approximation. In algebraic methods, μT is a local
function whereas in two equation model is a more general and complex functioned governing by two
additional PDEs. The second shortcoming is the need to make assumptions in evaluating the various
terms in model transport equation especially third order turbulent correlations19. The same short
coming plagues all other higher order closure attempts, so there is no magic bullet.

3.4.3.3 The RNG Model


The κ-ε model uses “highly abstruse‟ mathematics to extend the eddy viscosity turbulence model and
changes the governing equations by removing smaller scales of motion, replacing them with large
motions and modifying the viscosity term. Results have been good for backward-facing steps, while
other results have been mixed. The RNG model is not very commonly used and has a high
computational overhead.

3.4.3.4 The Realizable κ-ε Model


This is a non-linear version of the κ-ε model. It retains the two-equation κ-ε equations, but expands
the model by including additional effects to account for Reynold’s stress anisotropy without actually
using the seven extra equations used in the RSM (described later) to exactly model the Reynolds
stresses. The turbulent viscosity expression and the rate of dissipation of kinetic energy equation of
the standard κ-ε model are both changed in the realizable κ-ε model to take into account that
turbulence does not always adjust itself instantaneously while moving through the flow domain,
meaning that the Reynolds stress is partially dependent on the mean strain rate itself. This means
that the non-linear realizable κ-ε model allows for the phenomena of the state of turbulence lagging
behind the changes disturbing the turbulence production and dissipation balance.

3.4.3.5 Issues Relating to κ-ε Turbulence Modeling


Acceding to [Menter20], the κ-ε model has been very successful in a large variety of different flow
situations, but it also has a number of well-known shortcomings. From the standpoint of
aerodynamics, the most disturbing of them is the lack of sensitivity to adverse pressure-gradients.
and separation. Under those conditions, the model predicts significantly too high shear-stress levels
and thereby delays (or completely prevents) separation. This could be attributes this shortcoming to
the over prediction of the turbulent length-scale in the near wall region and has shown that a
correction proposed by [Hanjalic and Launder] improves the predictions considerably. However, the
correction is not coordinate-invariant and can therefore not be applied in general coordinates. An
alternative way of improving the results has been proposed by [Chen and Patel] and by [Rodi]. They
replace the ε-equation in the near wall region by a relation that specifies the length-scale analytically.
This also reduces some of the stiffness problems associated with the solution of the model. Although
the procedure is coordinate independent, it has only been applied to relatively simple geometries,

19 Anderson, Dale A; Tannehill, John C; Plecher Richard H; 1984,”Computational Fluid Mechanics and Heat
Transfer”, Hemisphere Publishing Corporation.
20 Florian R. Menter, “Improved Two-Equation k-ωTurbulence Models for Aerodynamic Flows”, NASA Technical

Memorandum 103975, October 1992.


29

where the change between the algebraic relation and the e-equation could be performed along a pre-
selected gridline. Clearly this cannot be done in flows around complex geometries. Furthermore, the
switch has to be performed in the logarithmic part (the algebraic length-scale is not known in the
wake region), so that the original k - ε model is still being used over most of the boundary layer.
Another problem with the k- ε model is associated with the numerical stiffness of the equations when
integrated through the viscous sublayer. This problem clearly depends on the specific version of the
k – ε model selected, but there are some general aspects to it. All low Reynolds number k – ε models
employ damping functions in one form or another in the sublayer. These are generally highly
nonlinear functions k of dimensionless groups of the dependent variables like Rt= κ2/εν (models
involving y+ are undesirable in separated flows). The behavior of these functions cannot easily be
controlled by conventional linearization techniques and can therefore interfere with the convergence
properties of the scheme. A second problem is that ε does not go to zero at a nonslip surface. There
is a significant number of alternative models that have been developed to overcome the shortcomings
of the κ-ε model. One of the most successful, with respect to both, accuracy and robustness, is the κ-
ω model of Wilcox21. It solves one equation for the turbulent kinetic energy k and a second equation
for the specific turbulent dissipation rate (or turbulence frequency) ω. The model performs
significantly better under adverse pressure-gradient conditions than the κ-ε model although it is the
authors experience that an even higher sensitivity to strong adverse pressure-gradients would be
desirable. Another strong-point of the model is the simplicity of its formulation in the viscous
sublayer. The model does not employ damping functions and has straightforward Dirichlet boundary
conditions. This leads to significant advantages in numerical stability.

3.4.3.6 The κ-ω and SST Models


As mentioned before, one of the main problems in turbulence modeling is the accurate prediction of
flow separation from a smooth surface. Standard two-equation turbulence models, such as κ-ε
models, often fail to predict the onset and the amount of flow separation under adverse pressure
gradient conditions. This is an important phenomenon in many technical applications, particularly
for airplane aerodynamics because the stall characteristics of a plane are controlled by the flow
separation from the wing. For this reason, the aerodynamic community has developed a number of
advanced turbulence models for this application. In general, turbulence models based on the ε
equation predict the onset of separation too late and under-predict the amount of separation later
on. This is problematic, as this behavior gives an overly optimistic performance characteristic for an
airfoil. The prediction is therefore not on the conservative side from an engineering stand-point. The
models developed to solve this problem have shown a significantly more accurate prediction of
separation in a number of test cases and in industrial applications. Separation prediction is important
in many technical applications both for internal and external flows. Currently, the most prominent
two-equation models in this area are the κ-ω based models of [Menter]22. The κ-ω based Shear-
Stress-Transport (SST) model was designed to give a highly accurate predictions of the onset and the
amount of flow separation under adverse pressure gradients by the inclusion of transport effects into
the formulation of the eddy-viscosity. This results in a major improvement in terms of flow
separation predictions.

21 W'llcox, D. C., "Reassessment of the Scale-Determining Equation for Advanced Turbulence Models," AIAA
Journal, Vol.26, Nov. 1988, pp.1299-1310.
22 Menter, F.R., “Two-equation eddy-viscosity turbulence models for engineering applications”, AIAA-Journal.,

32(8), pp. 1598 - 1605, 1994.


30

3.4.3.7 Wilcox (Standard) κ-ω Model


The starting point of the present formulation is the κ-ω model developed by [Wilcox]23. It solves two
transport equations, one for the turbulent kinetic energy, κ , and one for the turbulent frequency, ω.
The stress tensor is computed from the eddy-viscosity concept.

Dκ   ρκ  κ  ui
 ρP  β*ρωκ   μ  σ κ   with P  τ ij
Dt x j  ω  x j  xj
Eq. 3.22
Dω γω   ρκ  ω  ρσ d κ ω
 P - βρω 2   μ  σ ω  
Dt κ x j  ω  x j  ω x j x j

For recommendations for the values of the different parameters, see [Wilcox ]24.

3.4.3.8 The Baseline (BSL) κ-ω Model


The main problem with the Wilcox model is its well-known strong sensitivity to freestream
conditions [Menter]25. Depending on the value specified for ω at the inlet, a significant variation in
the results of the model can be obtained. This is undesirable and in order to solve the problem, a
blending between the κ-ω model near the surface and the κ-ε model in the outer region was
developed by [Menter]26. It consists of a transformation of the κ-ε model to a κ-ω formulation and a
subsequent addition of the corresponding equations. The Wilcox model is thereby multiplied by a
blending function F1 and the transformed κ-ε model by a function (1 - F1), F1 is equal to one near the
surface and decreases to a value of zero outside the boundary layer (that is, a function of the wall
distance). At the boundary layer edge and outside the boundary layer, the standard κ-ε model is
therefore recovered27.

3.4.3.9 SST κ-ω Model


The Shear-Stress Transport (SST) κ-ω turbulence model is a type of hybrid model, combining two
models in order to better calculate flow in the near-wall region. It was designed in response to the
problem of the κ-ε models unsatisfactory near-wall performance for boundary layers with adverse
pressure gradients. It utilizes a standard κ-ε model to calculate flow properties in the free-stream
(turbulent) flow region far from the wall, while using a modified k-ε model near the wall using the
turbulence frequency ω as a second variable instead of turbulent kinetic energy dissipation term ε.
Since the air in this projects heat exchanger is flowing between two flat fins very close to each other,
it is expected that the boundary layer flow has a strong influence on the results, and properly
modelling this near-wall flow could be important for accuracy of the calculations. Therefore the SST
κ-ω turbulence model has also been chosen for CFD simulations in this project. This SST κ-ω model
is similar to the κ-ε turbulence model, but instead of ε as the second variable, it uses a turbulence
frequency variable omega, which is expressed as ω = ε/k [s-1]. The SST κ-ω model computes Reynolds
stresses in the same way as in the κ-ε model. The transport equation for turbulent kinetic energy k
for the k-ω model is:

23 Wilcox, D. C. (2008), “Formulation of the k–ω Turbulence Model Revisited”, 46 (11), AIAA Journal, pp. 2823–
2838, Bibcode:2008AIAAJ..46.2823W, doi:10.2514/1.36541.
24 Wikipedia.
25 Menter, F.R., “Multiscale model for turbulent flows”, In 24th Fluid Dynamics Conference. American Institute of

Aeronautics and Astronautics, 1993.


26 Menter, F.R., “Two-equation eddy-viscosity turbulence models for engineering applications”, AIAA-Journal.,

32(8), pp. 1598 - 1605, 1994.


27 Training Manual, ANSYS CFX.
31

∂(ρκ) ∂ ∂ μT
+ (ρu̅i κ) = [(μ + ) ∇κ ] + P⏟κ − β⏟∗ ρκω
⏟∂t ∂xi
⏟ ∂xi
⏟ σκ
4 5
1 2 3
∂u̅i ∂u̅i 2 ∂u̅i
where Pκ = (2μT . − ρκ δ )
∂xj ∂xj 3 ∂xj ij
Eq. 3.23
The terms (1) - (5) in above expression, the turbulent kinetic energy k transport equation for the SST
Omega turbulence model, can be interpreted as the following:

1. Transient term - Accumulation of k (rate of change of k)


2. Convective transport - Transport of k by convection
3. Diffusive transport - Turbulent diffusion transport of k
4. Production term - Rate of production of k
5. Dissipation Rate of dissipation of k
Where σκ and β* are equation constants. The transport equation for turbulent frequency ω for the k-
ω model is:

∂(ρω) ∂(ρu̅i ω) ∂ μT
+ = [(μ + ) ∇ω] +
⏟∂t ⏟ ∂xi ∂x
⏟i σω,1
1 2 3
∂u̅i ∂u̅i 2 ∂u̅i ρ ∂κ ∂ω
γ2 (2ρ . − ρω δij ) − β⏟2 ρω2 + 2
⏟ ∂xj ∂xj 3 ∂xj σ
⏟ω,2 ∂xκ ∂xκ
5
4 6
Eq. 3.24
The general description for each of the terms in (1) to (6) are the usual terms for accumulation,
convection, diffusion, production, and dissipation of ω. The last term (6) is called a “cross-diffusion‟
term, an additional source term, and has a role in the transition of the modelling from ε to ω. The
constants for the Mentor SST κ-ω turbulence model are
listed in β* β2 σκ σω,1 σω,2 Υ2
Table 3.2. Additional modifications have been made to 0.09 0.083 1.0 2.0 1.17 0.44
the model for performance optimization. There are
blending functions added to improve the numerical Table 3.2 Constant values for SST κ-ω
stability and make a smoother transition between the turbulence model equations
two models. There have also been limiting functions
made to control the eddy viscosity in wake region and adverse pressure flows28. The κ-ω model is
similar to κ-ε, but it solves for ω, the specific rate of dissipation of kinetic energy. It also uses wall
functions and therefore has comparable memory requirements. Additionally, it has more difficulty
converging and is quite sensitive to the initial guess at the solution. Hence, the κ-ε model is often used
first to find an initial condition for solving the κ-ω model. The κ-ω model is useful in many cases
where the κ-ε model is not accurate, such as internal flows, flows that exhibit strong curvature,
separated flows, and jets29.

28 Versteeg, H K; Malalasekera, W.”An Introduction to Computational Fluid Dynamics, The Finite Volume
Method”, Second edition, Pearson Education Limited, Essex, England (2007).
29 Walter Frei, “Which Turbulence Model Should I Choose for My CFD Application? “, COMSOL Blog, September

16, 2013.
32

Case Study 1 – Subsonic Flow for Turbulence Models of (NACA)-0012 Airfoil


The analysis of the 2D subsonic flow over a (NACA) 0012 airfoil at various angles of attack and
operating at a Re = 3×106 is presented by [Eleni et al.]30. The flow was obtained by solving the steady-
state governing equations combined with one of three turbulence models (Spalart-Allmaras,
Realizable κ-ε, and κ-ω Shear Stress Transport (SST). The aim of the work was to show the behavior
of the airfoil at these conditions and to establish a verified solution method. The computational
domain was composed of 80000 cells emerged in a structured way, taking care of the refinement of
the grid near the airfoil in order to enclose the boundary layer approach. Calculations were done for
constant air velocity altering only the angle of attack for every turbulence model tested. This work
highlighted two areas in (CFD) that require further investigation: transition point prediction and
turbulence modeling. The laminar to turbulent transition point was modeled in order to get accurate
results for the drag coefficient at various Reynolds numbers. In addition, calculations showed that
the turbulence models used in commercial CFD codes does not give yet accurate results at high angles
of attack.

3.4.4.1 Computational Method


The NACA 0012, with Reynolds number of Re = 3x106, same with the reliable experimental data from
[Abbott and Von Doenhoff]31, in order to validate the present simulation. The free stream
temperature is 300 K, which is the same as the environmental temperature. The density of the air at
the given temperature is ρ=1.225kg/m3 and the viscosity is μ=1.7894×10-5 kg/ms. For this Reynolds
number, the flow can be described as incompressible. This is an assumption close to reality and it is
not necessary to resolve the energy equation. A segregated, implicit solver was utilized where the
calculations were done for angles of attack ranging from -12 to 20. The first step in performing a CFD
simulation should be to investigate the effect of the mesh size on the solution results. The appropriate
number of nodes can be determined by increasing the number of nodes until the mesh is sufficiently
fine so that further refinement does not change the results (mesh independence). Figure 3.4 shows
the effect of number of grid cells in coefficient of lift at stall angle of attack (16°). This study revealed
that a C-type grid topology with 80000 quadrilateral cells would be sufficient to establish a grid
independent solution. The
domain height was set to
approximately 20 chord
lengths, and the height of the
first cell adjacent to the
surface was set to 10-5,
corresponding to a maximum
y+ of approximately 0.2. A y+
of this size should be sufficient
to properly resolve the inner
parts of the boundary layer. In
order to include the transition
effects in the aerodynamic
coefficients calculation and get
accurate results for the drag
coefficient, a new method was
used. The transition point Figure 3.4 Lift Coefficient at Stall (AoA) against Number of Cells

30 D. C. Eleni, Tsavalos I. Athanasios and Margaris P. Dionissios, “Evaluation of the turbulence models for the
simulation of the flow over a National Advisory Committee for Aeronautics (NACA) 0012 airfoil”, Journal of
Mechanical Engineering Research, March 2012.
31 Abbott IH, Von Doenhoff AE. “Theory of Wing Sections”. Dover Publishing, New York.
33

from laminar to turbulent flow on the airfoil was determined and the computational mesh was split
in two regions, a laminar and a turbulent region. To calculate the transition point the following
procedure was used. A random value for the transition point (xtr) was chosen and the computational
domain was split at that point with a perpendicular line. The problem was simulated by defining the
left region as laminar and the right as turbulent zone.

3.4.4.2 Results and Discussion


Simulations for various angles of attack were done in order to be able to compare the results from
the different turbulence models and then validate them with existing experimental data from reliable
sources. To do so, the model was solved with a range of different angles of attack from -12 to 20°. On
an airfoil, the resultants of the forces are usually resolved into two forces and one moment. The
component of the net force acting normal to the incoming flow stream is known as the lift force and
the component of the net force acting parallel to the incoming flow stream is known as the drag force.
The curves of the lift and the drag coefficient are shown for various angles of attack, computed with
three turbulence models and compared with experimental data. Figure 3.5 shows that at low angles
of attack, the dimensionless lift coefficient increased linearly with angle of attack. Flow was attached
to the airfoil throughout this regime. At an angle of attack of roughly 15 to 16°, the flow on the upper
surface of the airfoil began to separate and a condition known as stall began to develop. All three
models had a good agreement with the experimental data at angles of attack from -10 to 10° and the
same behavior at all angles of attack until stall. It was obvious that the Spalart-Allmaras turbulence
model had the same behavior with the experimental data as well as after stall angle.

Figure 3.5 Comparison Between Experimental Data [Abbott et al.] and Three Different
Turbulent Models Simulation Results for Lift Coefficient in NACA 0012 Airfoil
34

Near stall, disagreement between the data was shown. The lift coefficient peaked and the drag
coefficient increased as stall increased. The predicted drag coefficients were higher than the
experimental data (Figure 3.6). This over prediction of drag was expected since the actual airfoil
has laminar flow over the forward half. The turbulence models cannot calculate the transition point
from laminar to turbulent and consider that the boundary layer is turbulent throughout its length.
From theory, the turbulent boundary layer carries more energy and is much greater than at the
viscous boundary layer, which carries less energy. The computational results must be compared with
experimental data of a fully turbulent boundary layer. This was done only for CD as CL is less sensitive
to the transition point.
[ Johansen]32 contained experimental data of CD for the NACA 0012 airfoil and Re = 3×106, where the
boundary layer formed around the airfoil is fully turbulent. Figure 3.6 shows the curves of CD for
various angles of attack, compared with experimental data for fully turbulent boundary layer33. The
values of from the three turbulence models were very close to experimental data for the fully
turbulent boundary layer. The most accurate model was the κ-ω SST model, next came the Spalart-
Allmaras, and latest in precision was the Realizable κ-ε .
In order to get more accurate results, the computational domain could be split into two different
domains to run mixed laminar and turbulent flow. The disadvantages of this approach were that the
accuracy of simulations depends on the ability to accurately guess the transition location, and a new
grid must be generated if the transition point had to change [Silisteanu-Botez]34. If the transition
point is known, the grid can easily be split in two with a vertical line that passes through this point
and then laminar and turbulent zones
are defined. The results of this method
at angle of attack a=0 and operating at
Re = 1×106, 2×106, 3×106, 4×106 and
5×106. Initially, was calculated for a
fully turbulent boundary layer and
compared with CD experimental data
from NASA [McCroskey]35. Then,
simulations were made with the split
grid for the five Reynolds numbers. The
computational results for the fully
turbulent boundary layer agreed very
well with the corresponding
experimental data. The discrepancy
between the Drag Coefficient and
experimental data from [McCroskey]
for fully turbulent boundary layer was
up to 5.6%. On the other hand, the
comparison between the simulation
Figure 3.6 Comparison Between Experimental Data for
results with the split grid and the
Transitional Boundary Layer and Different Turbulent Models
experimental data from [McCroskey] on the Drag Coefficient of NACA-0012 Airfoil

32 Johansen J. “Prediction of Laminar/Turbulent Transition in Airfoil Flows”. RISE National Laboratory, Roskilde,
Denmark, 1997.
33 Comparison Between Different Turbulent Models and Experimental Data obtained by [Abbott & Von

Doenhoff] and [Johansen] for Transitional Boundary Layer on the Drag Coefficient of NACA-0012 Airfoil.
34 Silisteanu PD, Botez RM. “Transition flow occurrence estimation new method”. 48th AIAA Aerospace Science

Meeting. Orlando, Florida, 2010.


35 McCroskey WJ, ”A Critical Assessment of Wind Tunnel Results for the NACA 0012 Airfoil”. U.S. Army Aviation

Research and Technology Activity, Nasa Technical Memorandum, 42: 285-330, 1987.
35

for transitional boundary layer


showed an excellent agreement, with
maximum error of about 3.6%. It was
also observed that as the Reynolds
number increased, the Drag
Coefficient decreased. When the
boundary layer was fully turbulent
the reduction of was more intense
and when there was a transition from
laminar to turbulent, was reduced to
a much lower rate. It is worth noting
that this process of calculating the
transition point is quite simple when
the angle of attack is zero because the
flow is symmetric and the transition
point is the same above and below
the airfoil. At nonzero angles of
attack the process is more
complicated because transition
points are different for the upper and
lower surface of the airfoil. With
different AoA, the pressure on the
lower surface of the airfoil was
greater than that of the incoming
flow stream and as a result it
effectively “pushed” the airfoil
upward, normal to the incoming flow
stream. On the other hand, the
components of the pressure Figure 3.7 Contours of velocity magnitude at 9° (Top) and 16°
distribution parallel to the incoming (Bottom) AoA with the Spalart-Allmaras turbulence model
flow stream tended to slow the
velocity of the incoming flow relative
to the airfoil, as do the viscous stresses. Contours of velocity components at angles of attack 9 and
16° are also shown in Figure 3.7. The trailing edge stagnation point moved slightly forward on the
airfoil at low angles of attack and it jumped rapidly to leading edge at stall angle. A stagnation point
is a point in a flow field where the local velocity of the fluid is zero. The upper surface of the airfoil
experienced a higher velocity compared to the lower surface. That was expected from the pressure
distribution. As the angle of attack increased the upper surface velocity was much higher than the
velocity of the lower surface. For further information, please consult the [Eleni et al.]36.

3.4.4.3 Conclusions
This paper showed the behavior of the 4-digit symmetric airfoil NACA 0012 at various angles of
attack. The most appropriate turbulence model for these simulations was the κ-ω SST two-equation
model, which had a good agreement with the published experimental data of other investigators for
a wider range of angles of attack. The predicted drag coefficients were higher than the existing
experimental data from reliable sources. This over prediction of drag was expected since the actual

36D. C. Eleni, Tsavalos I. Athanasios and Margaris P. Dionissios, “Evaluation of the turbulence models for the
simulation of the flow over a National Advisory Committee for Aeronautics (NACA) 0012 airfoil”, Journal of
Mechanical Engineering Research, March 2012.
36

airfoil has laminar flow over the forward half. The computational results from the three turbulence
models were compared with experimental data where the boundary layer formed around the airfoil
is fully turbulent and they agreed well. Afterwards, the transition point from laminar to turbulent
regime was predicted, the computational grid split in two regions, a laminar and a turbulent region,
and then new simulations were realized. By this method, the computational results agreed very well
with corresponding experimental data37.

3.5 Near Wall Turbulence Modeling


Most turbulence models requires special algebraic formula, often called Wall Function to represent
the distribution of velocity, temperature, turbulence, energy, etc. within boundary layers. This
practice is necessary because these models are not valid in the region within the layer where the
molecular and turbulent effects are comparable in magnitude. It is also expedient because it avoids
the need for employing a fine mesh within boundary layer. The wall function representation of near-
wall turbulent behavior is inexact, the accuracy being dependent on the degree to which the
assumption and approximations embodied in the function correspondent with reality of the
application. The Standard wall function with following characteristics:

 Variation in velocity etc. are predominantly normal to the wall, leading to one dimensional
behavior.
 Effects of pressure gradients and body forces are negligible, leading to uniform shear stress
in the layer.
 Shear stress and velocity vectors are aligned and unidirectional through the layer.
 A balance exist between turbulence energy production and dissipation.

Figure 3.8 Zones in Turbulent B. L. for a typical Incompressible flow over a smooth flat plate

37 See Previous.
37

 There is a linear variation of turbulence length scales.

For flow along solid surfaces (internal or external), a good approximation is observed by evaluating
the expressions for linner is responsible for producing the inner law of the wall region of turbulent flow
and louter produces the outer wake-like region. These two zones are illustrated in Figure 3.8 which
depicts a typical velocity distribution for an incompressible turbulent boundary layer on a smooth
impermeable plate.

Modeling Flow Near the Wall


Two approaches are commonly used to model the flow in the near-wall region:

3.5.1.1 Wall Function Method


uses empirical formulas that impose suitable conditions near the wall without resolving the
boundary layer, thus saving computational resources. All turbulence models address a suitable wall
function method. The major advantages of the wall function approach is that the high gradient shear
layers near walls can be modeled with relatively coarse meshes, yielding substantial savings in CPU
time and storage. It also avoids the need to account for viscous effects in the turbulence model. (see
Figure 3.9).

3.5.1.2 Low-Reynolds-Number Method


Resolves the details of the boundary layer profile by using very small mesh length scales in the
direction normal to the wall (very thin inflation layers). Note that the low-Re method does not refer
to the device Reynolds number, but to the turbulent Reynolds number, which is low in the viscous
sublayer. This method can therefore be used even in simulations with very high device Reynolds
numbers, as long as the viscous sublayer has been resolved. The computations are extended through
the viscosity-affected sublayer close to the wall. The low-Re approach requires a very fine mesh in
the near-wall zone and correspondingly large number of nodes. Computer-storage and run-time
requirements are higher than those of the wall-function approach and care must be taken to ensure
good numerical resolution in the near-wall region to capture the rapid variation in variables. To
reduce the resolution requirements, an automatic wall treatment was developed which allows a
gradual switch between wall functions and low-Reynolds number grids, without a loss in accuracy.

Figure 3.9 Turbulence and near Wall Function


38

Case Study 2 - Wall-Modelling Strategies in Large Eddy Simulation of Separated High-


Reynolds-Number Flows
If ReL (say, that based on the boundary layer thickness) exceeds roughly 105, the resource
requirements are entirely dominated by the need to resolve the near-wall region (the inner layer),
with the number of nodes rising roughly as N ~ Re2.5L. Simulations for practical configurations would
not, in most circumstances, be undertaken today with meshes exceeding 10–50M nodes,
corresponding to ReL ~ 5 x 105, yet this is still a very modest Reynolds number in practice [Tessicini,
and Leschziner]38. The general approach taken in recent years towards addressing the problem of
near-wall resolution has been to combine RANS modelling near the wall with LES in the outer flow.
The key premise underpinning this strategy is that it should allow near-wall numerical cells to be
used that have far higher aspect ratios than those required by LES, typically 500-1000, relative to 50
in wall-resolving LES. Substantial savings could thus be made by using much coarser stream wise and
span wise meshes than are dictated by the LES constraints. Efforts currently have focused on two
particular methods, one being a hybrid LES-RANS scheme and the other being a two-layer zonal
schemes. The difference between them is explained by reference to Figure 3.10, which conveys the
manner by which the LES and RANS regions communicate numerically.
The hybrid method uses a single computational domain. Within a predefined layer near the wall, that
can be prescribed in terms of y+, RANS equations are solved using one-equation or two-equation
eddy-viscosity models that are dynamically adjusted so to comply with continuity of eddy viscosity
across the interface, νRANS= νLES, beyond which a LES sub-grid-scale model is used. To achieve this
compatibility, the RANS model coefficients at the interface is determined by comparison of the RANS
viscosity, containing the RANS-determined turbulence energy and dissipation rate at the interface,
to the LES viscosity at the interface. The variation of the coefficients from the interface to the wall is
then prescribed analytically, based on observations derived from a-priori wall-resolved LES

Figure 3.10 Schematics of hybrid LES-RANS scheme (upper) and two-layer zonal scheme (lower)

38F. Tessicini, M.A. Leschziner, “Wall-modelling strategies in large eddy simulation of separated high- Reynolds-
number flows”.
39

performed in channel flows. The zonal method uses two overlapping grids across the near-wall layer.
The LES grid extends to the wall, but is relatively coarse, maintaining cell-aspect-ratio constraints
appropriate to LES. Within the near-wall layer, a separate grid is inserted, which is refined towards
the wall, typically to a wall-nearest node located at y+=O(1). Within that layer, parabolized RANS
equations are solved for the wall-parallel-velocity components, using a simple algebraic turbulence
model - for example, a mixing-length model.

3.6 Case Study 3 - Numerical Modeling of Turbulence in Turbine Stages


Many of existing turbulence models are applied for modeling and research of turbulence effects in
turbine stages [lieva]39. Due to the strong accelerations and decelerations of flow in turbine cascades,
the local value of free‐stream turbulence, at the location of boundary layer transition onset, may
significantly vary from first to the last turbine stage. Currently applied transition onset correlations
involve data from many scientists, who have adopted different approaches to define the free‐stream
turbulence values. Laminar separation bubbles can result from laminar separation followed by
sufficiently early transition in the separated shear layer and subsequent turbulent reattachment.
Errors in predicting the length of these bubbles will lead to failures in the blade design and wrong
solutions. aryl attempts at describing bubble development and separation, perceive40-41, were based
on semi‐empirical models. In those models, constant pressure for the region of the separated laminar
shear layer, instantaneous transition, and linear variations in free‐stream velocities during the phase
of turbulent reattachment was assumed. An integral boundary layer computation procedure was
applied, the location of the transition onset was found with correlations for separated laminar layer,
in function of the momentum thickness and Re.

Background and Literature Survey


Flows with transition and separation phenomena could be modeled with application of
contemporary models for accurate description of all aerodynamic effects, which are expected to be
observed, together with innovative and very correct transition model. Modeling of bubble dynamics
is important for the purposes of research and prediction of separated flows with transition42.
According to43-44 various predictive techniques were described in detail in the area of turbulence in
turbines and stressed on the need for application of improved and correct transition modeling. In the
literature, many papers discuss experiments and their outcomes related to turbine blades. Mostly,
they present research in more global aspect, a small number of experiments provide detailed results
useful for turbulence modeling. This is related to the fact that it is very difficult to obtain sufficiently
thick boundary layers to perform detailed measurements on the suction and pressure sides. Many
results are obtained after modeling and measurements related to research on a flat plate with a

39 Galina Ilieva Ilieva, “On Turbulence and its


Effects on Aerodynamics of Flow through Turbine Stages”, Technical
University, Varna, Bulgaria, Intech Open 2017. http://dx.doi.org/10.5772/intechopen.68205
40 Horton HP. A Semi‐Empirical Theory for the Growth and Bursting of Laminar Separation Bubbles. 1969.
41 Roberts WB. Calculation of laminar separation bubbles and their effects on airfoil performances. AIAA Journal.

1980;18: 25‐31.
42 Jahanmiri M. Boundary Layer Transitional Flow in Gas Turbines, Research Report 2011:01, Göteborg, Sweden:

Division of Fluid Dynamics Department of Applied Mechanics, Chalmers University of Technology; 2011.
43 Lakshminarayana B. “An assessment of computational fluid dynamic techniques in the analysis and design of

turbomachinery”, ASME Journal of Fluids Engineering. 1991;113: 315‐352.


44 Yaras MI. Measurements of the Effects of Freestream Turbulence on Separation Bubble Transition. ASME Paper

No. GT‐2002‐30232; 2002.


40

pressure gradient, imposed by the external wall45-46 for negative pressure gradients; also, after
application of Görtler vortex on the concave plate47. Results for the streaming effects of blade convex
side, are shown in48. Studies in49 discuss results of measurements on the suction surface of blade
under conditions of very low Reynolds number.
There are mainly two approaches used to model bypass transition in industry50. The first is to apply
low‐Reynolds number turbulence models in which wall‐damping functions implemented into the
turbulent transport equations were applied to obtain the moment when boundary layer transition
will occur. Research activities have proved that this approach cannot predict very well the influence
of various factors, such as pressure gradients, free‐stream turbulence, and wall roughness to predict
the transition onset. Damping functions, optimized to damp the turbulence in the viscous sublayer,
cannot give reliable prediction of the transition when subjected different and complicated processes
51.

The second approach is application of experimental correlations related to the free‐stream


turbulence intensity and to the transition Reynolds number, with included boundary layer
momentum thickness52,53. The last approach is proved as accurate, but very challenging, actual
momentum‐thickness, Reynolds numbers must be compared with their critical values, obtained
from correlations, included into the mathematical model. Applications purposes to arrive to
physically correct numerical solution. There are additional difficulties related to application of
unstructured mesh, not well‐defined boundary layer, and various approaches to attain numerical
solution54.

Transition Phenomenon as Relates to Gas Turbines


Speaking about transition and its effects on the entire flow field, it is necessary to mention that flow
passing through a turbine stage is essentially turbulent and unsteady. The nonstationary pulsations
are obtained as a result from the stator‐rotor interaction, mainly. Periodic phenomenon, caused by
stator‐rotor interaction effects, excites both the flow, passing over blade surfaces, and boundary layer
characteristics. This results in an increased production of the so‐called turbulent spots and shifts the
location of laminar‐turbulent transition in the upstream direction. This laminar‐turbulent transition
phenomenon is known as “wake‐induced transition”.

45 Keller FJ, Wang T. Flow and Heat Transfer Behavior in Transitional Boundary Layers with Stream wise
Acceleration. ASME 94‐GT‐24; 1994.
46 Kestoras MD, Simon TW. Effect of Free‐Stream Turbulence Intensity on a Boundary Layer Recovering from

Concave Curvature Effects. ASME 93‐GT‐25; 1993.


47 Peerhossaini, H, Wesfreid JE. Experimental study of the Görtler‐Taylor instability. Wesfreid JE, Brand H,

Mannville P, Albinet G, Boccara N, editors. Propagation in Systems for From Equilibrium. Berlin: Springer; 1988.
48 Sharma OP, Wells RA, Schlinker RH, Bailey DA. Boundary layer development on turbine airfoil suction surfaces.

ASME Journal of Engineering for Power 1982;104:698‐706.


49 van Treuren KW, Simon T, von Koller M, Byerley AR, Baughn JW, Rivir R. Measurements in a turbine cascade

flow under ultra-low Reynolds number conditions. Journal of Turbomachinery, Transaction of ASME.
2002;124:100‐106.
50 Mayle RE. The role of laminar‐turbulent transition in gas turbine engines. ASME Journal of Turbomachinery.

1991;113: 509‐537.
51 Reynolds O. An experimental investigation of the circumstances which determine whether the motion of water

shall be direct or sinuous, and of the law of resistance in parallel channels. Philosophical Transactions of the Royal
Society of London A. 1883;174:935‐982.
5252 Mayle RE. The role of laminar‐turbulent transition in gas turbine engines. ASME Journal of Turbomachinery.

1991;113: 509‐537.
53 Abu‐Ghannam BJ, Shaw R. Natural transition of boundary layers‐the effects of turbulence, pressure gradient,

and flow history. Journal of Mechanical Engineering Science. 1980;22(5):213‐228.


54 Reza Taghavi Z, Salary M, Kolaei A. Prediction of boundary layer transition based on modeling of laminar

fluctuations using RANS approach. Chinese Journal of Aeronautics. 2009;22:113‐120.


41

Figure 3.11 depicts a picture of possible boundary layer development over surfaces of high pressure
blade is shown. On the suction side, it is usually expected that in the downstream direction of the
initial laminar part, a boundary layer will transfer to turbulent (2) in Figure 3.11. The size of the
transition zone is related to the place where transition phenomenon could be observed ‐ in upstream
direction or downstream direction of the place of minimum pressure. In the upstream direction, the
zone of transition is
expected to comprehend
bigger area. If a laminar
separation bubble occurs
in the front part of the
suction side (1), then the
presence of high
pressure gradients will
force the boundary layer
to develop as laminar
again in downstream
direction; forward
transition will take place.
The reverse transition
may appear on the
suction surface. In the
case of research on film‐
cooled gas turbine
blades, the transition is
expected to appear at the Figure 3.11 Boundary Layer Development on High Pressure Turbine
places where cooling jets Blades (Mayle)
are injected in the main
flow.
In downstream, a reverse transition process also could be recognized. This fact could affect the heat
transfer distribution over surfaces of film‐cooled blades. On the profile pressure side ‐ if a separation
bubble occurs, the reattached turbulent boundary layer may become again laminar like, (2) in Figure
3.11. In the case of lack of separation bubble, a forward transition zone, followed by a reverse one,
in the rear part of the profile, could be observed (1) in Figure 3.11.
In high pressure turbines, the effect of transition on losses is usually small, because the aerodynamic
42

losses are mainly related to the turbulent flow development after the moment of transition. In low
pressure turbines, the
flow in inter blade
channels is characterized
by low Re. Especially for
gas turbines, as part of
aircraft engines, the
operating Reynolds
numbers are low at high
altitudes to begin with
and a further decrease
can cause separation
before transition.
In regions where
expansion occurs, the
fluid is highly
accelerated and the
boundary layer has small
thickness due to the
Figure 3.12 Transition on a Low‐Pressure Turbine Airfoil at Various
favorable pressure
Reynolds Number (Courtesy of lieva)
gradients. At high
Reynolds numbers,
transition occurs far in the upstream direction, flow is mainly turbulent over the profile. Near the
trailing edge, in function of the blade profile geometry, the boundary layer will separate forced by
turbulent levels. When Re number decreases, turbulent separation disappears and transition (the
“bypass transition”) moves in the downstream direction; at that moment losses are minimal. If Re
number decreases more, laminar separation ahead of the transition region could appear. In the case
of no separation, the bubble is small enough so that the flow could reattach to the blade surface. In
this case, aero-dynamical losses are slightly higher than the previously described case. For lower
Reynolds numbers the increase of laminar shear layer and transition length, until reattachment,
before the trailing edge, is no longer possible and thus a complete separation occurs, see Figure
3.12.

Example of Modeling of Turbine Stage with Twisted Rotor Blade with Different Turbulence
Models
The main target of this research in which geometry modeling, numerical set‐up, and convergence
problem solution are described as detailed in - , is to define the flow parameters distribution in a 3D
turbine stage with twisted rotor blade. For the purposes of the turbulence modeling, the standard k
– ε turbulence model, RNG k – ε, standard k – ε, for the case of research on radial gap, and RSM
(Reynolds stress model) models are applied, in regard to the flow conditions. The Reynolds stress
model (RSM) is applicable for modeling effects of additional vortices, found in flow and shear stress
effects over fluid particles . The standard k – ε model gives quite good values, especially for the
turbulent kinetic energy, in the core flow see. In , results show that the advantage of using the RSM
in regions of flow separation; however, the main flow features were still good enough, captured by
the k – ε model. The RNG model gives the highest prediction of lift and maximal lift angle , . The k –
ε turbulence models are appropriate for flows characterized by high adverse pressure and intensive
separation. This model allows for a more accurate near wall treatment with an automatic switch from
wall function to low‐Reynolds number formulation, based on grid spacing.
43

Results
In the current study, it is found that depending on the specific flow feature, under consideration,
different turbulence model have to be applied. Numerical results for pressure distribution, in the
case of applied standard k – ε model, are shown in Figure 3.13. For visualized vortices, in radial
direction, due to difference between the pressure field values for hub and shroud sections in the
turbine stage, in the case of applied Reynolds stress model (RSM), see [lieva]55. The area occupied by
this vortex is bigger than the one formed in the case of standard k – ε turbulence model, see [lieva]56.
The outcomes of the performed research are as follows:

 The RNG model is acceptable to study both the shear stress and streamlines curvature effects.
It presents vortices formed at the trailing edge and also provides results for aerodynamic
features at the leading edge.

 In the case of applied RSM model, a relative decrease of 1.308% for turbine stage efficiency
is observed. This is a result of taking into account of all pulsations and vortex structures near
the wall regions, boundary layer separation, viscosity, and compressibility effects.

 The RNG k – ε turbulence model leads to increased values for turbulent intensity and less
turbulent viscosity. This is a prerequisite for decrease of the left‐hand side term values in the
momentum equations, furthermore causes relative increase of stage efficiency with 0.147%,
in a comparison with the case of implemented RSM turbulence model.

Figure 3.13 Static Pressure Distribution (Courtesy of lieva)

55 Galina Ilieva Ilieva, “On Turbulence and its Effects on Aerodynamics of Flow through Turbine Stages”, Technical
University, Varna, Bulgaria, Intech Open 2017. http://dx.doi.org/10.5772/intechopen.68205
56 Galina Ilieva Ilieva, “On Turbulence and its Effects on Aerodynamics of Flow through Turbine Stages”, Technical

University, Varna, Bulgaria, Intech Open 2017. http://dx.doi.org/10.5772/intechopen.68205


44
45

4 Beyond the Boussinesq Approximation (Non-Linear Modeling)


The Boussinesq eddy-viscosity approximation assumes the principal axes of the Reynolds-stress
tensor, τij, are coincident with those of the mean strain-rate tensor, Sij, at all points in a turbulent flow.
This is the analog of Stokes approximation for laminar flows. The coefficient of proportionality
between τij and Sij is the eddy viscosity, μT, which is linear. Unlike the molecular viscosity which is a
property of the fluid, the eddy viscosity depends upon many details of the flow under consideration.
It is affected by the shape and nature (e.g., roughness height) of any solid boundaries, freestream
turbulence intensity, and, perhaps most significantly, flow history effects. Experimental evidence
indicates that flow history effects on τij often persist for long distances in a turbulent flow, thus
casting doubt on the validity of a simple linear relationship between τij and Sij. Next, we outline
several flows for which the Boussinesq approximation yields a completely unsatisfactory description.

4.1 Boussinesq-Approximation Deficiencies


While models based on the Boussinesq eddy-viscosity approximation provide excellent predictions
for many flows of engineering interest, there are some applications for which predicted flow
properties differ greatly from corresponding measurements. Generally speaking, such models are
inaccurate for flows with sudden changes in mean strain rate and for flows with what Bradshaw
refers to as extra rates of strain. It is unsurprising that flows with sudden changes in mean strain rate
pose a problem. The Reynolds stresses adjust to such changes at a rate unrelated to mean flow
processes and time scales, so that the Boussinesq approximation must fail. Similarly, when a flow
experiences extra rates of strain caused by rapid dilatation, out of plane straining, or significant
streamline curvature, all of which give rise to unequal normal Reynolds stresses, the approximation
again becomes suspect. Some of the most noteworthy types of applications for which models based
on the Boussinesq approximation fail are:

1. flows with sudden changes in mean strain rate;


2. flow over curved surfaces;
3. flow in ducts with secondary motions;
4. flow in rotating and stratified fluids ;
5. three-dimensional flows;
6. flows with boundary-layer separation .

4.2 Nonlinear Constitutive Relations


One approach to achieving a more appropriate description of the Reynold stress tensor without
introducing any additional differential equations is to assume the Boussinesq approximation is
simply the leading term in a series expansion of functional. Proceeding with this premise it can be
shown that for incompressible flow the expansion must proceed through second order according to

 ui u j  2
 ρu i uj  μ T     δ ij ρ κ  or
 x x  3
 j i 
Eq. 4.1
2 1  u u j 
 ρu i uj   TSij  δ ij κ where Sij   i  
3 2  x j x i 

The above isotropic relation assumes that the principal axis of the Reynolds stress tensor S͞ ij coincides
with that of the mean strain rate. The standard κ-ε model does not take into account the anisotropic
effects and fails to represent the complex interaction mechanisms between Reynolds stresses and
46

the mean velocity field. For example, the linear model fails to mimic the effects related to streamline
curvature, secondary motion, or flow with extra strain rates. These anisotropic effects can be
predicted by introducing a nonlinear expression for the Reynolds stresses as given in the following
expression57:

2 ρ̅κ ρ̅κ
̅̅̅̅̅̅̅
−ρu ′ u′ = − ρ
i j ̅ κδij + 2μT S̅ij − B 2 S̅mn S̅nm δij − C 2 Sik Skj −
⏟3 ⏟ω ω
Boussinesq Non−Linear Trem 1
ρ̅κ ρ̅κ ρ̅κ
D 2 (S̅ik Ω
̅ kj + S̅jk Ω
̅ ki ) − F ̅ mn Ω
Ω ̅ nm δij − G ̅ Ω
Ω ̅
⏟ω ω2 ω2 ik kj
Non−Linear Term 2
κ κ3 1 ∂u̅ ∂u̅ 1 ∂u̅ ∂u̅
where ≈ , ̅ ij = ( i + j )
S ̅ ij = ( i − j )
and Ω
ω 2 ε2 2 ∂xj ∂xi 2 ∂xj ∂xi
Eq. 4.2
Where B, C, D, F, and G are closure coefficients. These coefficients of the non-linear terms should be
carefully determined because they are expected to influence the physical accuracy and numerical
performance of the model. Here, the coefficients are adjusted through the consideration of the
anisotropy in simple shear flows detailed by [Champagne et al.]58 and [Harris et al.]59 Solving this
equation is a daunting task, but there are assumptions (as the case with all turbulence models) which
can be made, depending to the case. We don’t get into different modeling but for an excellent
discussion, readers should refer to60. (See Eq. 4.2).

Case Study – 3D Simulation of Flow Past a Cylinder using Nonlinear Turbulence Model
The flow past a cylinder of circular cross section has been the subject of interest for industrial
researchers as well as scientists, because of its wide range of applications. To cite a few examples,
flow in bridge piers, chimney stacks, and tower structures in civil engineering; electrodes in chemical
engineering; nuclear fuel rods in the atomic field and heat exchanger tubes in thermal engineering,
etc., fall under this subject of study. Although the geometry is simple, the flow has complicated
features such as stagnation points, laminar boundary-layer separation, turbulent shear layers,
periodic vortex shedding, and wakes. Even though there is much literature available on numerical
simulation of laminar flow past a two-dimensional circular cylinder at low Reynolds number, a focus
on practical high Reynolds numbers is less. This could be due to the complexity of formulating
Reynolds stresses in turbulent flows. The majority of turbulent flow calculations carried out in earlier
days used two equation models such as the standard κ-ε model (hereafter referred as SKE) and the
κ-ω model, because of their robustness, computational efficiency, and completeness. In the classical
SKE model, the turbulent kinetic energy (κ) and the turbulent kinetic energy dissipation rate (ε) were
calculated using modeled transport equations separately for k and e along with the Boussinesq eddy
viscosity approximation.

57 Ichiro Kimura, and Takashi Hosoda, ”A non-linear κ-ε model with realizability for prediction of flows around
bluff bodies”, Int. J. Numerical Meth. Fluids 2003; 42:813–837 (DOI: 10.1002/_d.540).
58 Champagne FH, Harris VG, Corrsin S. ,”Experiments on nearly homogeneous turbulent shear flow”. Journal of

Fluid Mechanics 1970; 41:81–139.


59 Harris VG, Graham JAH, Corrsin S. “Further experiments in nearly homogeneous turbulent shear flow”. Journal

of Fluid Mechanics 1977; 81:657–687.


60 David C. Wilcox, ”Turbulence Modeling for CFD”,1993, 1994 by DCW Industries, Inc.
47

4.2.1.1 Literature Survey


Subsequently, nonlinear models were proposed by many researchers such as [Gatski and Speziale]61,
[Craft et al].62, and Shih et al63. Some researchers have tested the appropriateness of nonlinear
turbulence models for simple flows such as homogeneous shear flow, separated flow over a
backward-facing step, and flow in a confined jet. They obtained encouraging results in the prediction
of mean and turbulent quantities. The two-dimensional, unsteady, Reynolds-averaged Navier-Stokes
(URANS) simulation of subcritical flow past a circular cylinder at Re = 1.4x105 was carried out by
[Tutar and Holdo]64 using the SKE, nonlinear κ-ε formulation and large-eddy simulation (LES)
technique. The wake centerline velocity recovery predicted by the nonlinear κ-ε turbulence model
showed better agreement with the LES and experimental results than those predicted by the SKE
model. [Saghafian et al.]65 simulated the two dimensional flow past a circular cylinder using a
nonlinear eddy viscosity model and tested the mean drag coefficient (Cd mean), root-mean-square
(RMS) lift coefficient (Cl RMS), and Strouhal number value (St = f D/U∞) for the range of Reynolds
numbers from subcritical laminar separation to supercritical turbulent separation.
The standard and renormalized group versions of the κ-ε model were examined by Jennifer 66 for
two-dimensional flow past a circular cylinder at subcritical Reynolds number Re=5,232. With this
increasing interest in nonlinear turbulence models, [Kimura and Hosoda]67 proposed a cubic
nonlinear κ-ε model by accounting for the effect of anisotropy. They tested the model for two-
dimensional flow around a square cylinder and a surface-mounted cubic obstacle. The model, which
included the realizability condition, performed better than linear models when compared with
experimental results. Recently, the above model was used by [Ramesh et al.]68 for simulating three
dimensional flow around a square cylinder, and they found it performs better than results by the
standard κ-ε and RNG κ-ε models.

4.2.1.2 Basic Equations


The ensemble-averaged RANS equations for an incompressible flow are Continuity and Momentum
equations given by:

u i u i u i u j P (ui uj )  2ui


0 ,     Re 1 Eq. 4.3
x i t i x j x i x j x i x j
where xi is the spatial coordinate, t is the time, u͞i is the ensemble-averaged velocity, ͞u´i is the

61 T. B. Gatski and C. G.Speziale, “On Explicit Algebraic Stress Models for Complex Turbulent Flows”, J. Fluid Mech.,
vol. 254, pp. 59–78, 1993.
62 T. J. Craft, B. E. Launder, and K. Suga, “Development and Applications of a Cubic Eddy-Viscosity Model of

Turbulence”, Int. J. Heat Fluid flow, vol. 17, pp. 108–115, 1996.
63 T. H. Shih, J. Zhu, and J. L. Lumley, “A New Reynolds Stress Algebraic Equation Model”, Computer. Meth. Appl.

Mech. Eng., vol. 125, pp. 287–302, 1997.


64 M. Tutar and A. E. Holdo, “Computational Modeling of Flow around a Circular Cylinder in Subcritical Flow

Regime with Various Turbulence Models”, Int. J. Numerical Methods Fluids, vol. 35, pp. 763–784, 2000.
65 M. Saghafian, P. K. Stansby, M. S. Saidi, and D. D. Asplay, “Simulation of Turbulent Flows around a Circular

Cylinder Using Non-Linear Eddy-Viscosity Modeling: Steady and Oscillatory Ambient Flows”, J. Fluids Structure,
vol. 17, pp. 1213–1236, 2003.
66 R. B. Jennifer, “Verification Testing in Computational Fluid Dynamics: An Example Using Reynolds-Averaged

Navier-Stokes Methods for Two Dimensional Flow in the Near Wake of a Circular Cylinder”, Int. J. Numerical Meth.
Fluids, vol. 43, pp. 1371–1389, 2003.
67 L. Kimura and T. Hosoda, “A Non-linear κ-ε Model with Reliability for Prediction of Flows around Bluff Bodies”,

Int. J. Numerical Method Fluids, vol. 42, pp. 817–837, 2003.


68 V. Ramesh, S. Vengadesan, and J. L. Narasimhan, “3D Unsteady RANS Simulation of Turbulent Flow over Bluff

Body by Non-linear Model”, Int. J. Numerical Meth. Heat Fluid Flow, vol. 16, pp. 660–673, 2006.
48

fluctuating velocity, and P is the averaged pressure divided by the density. As a result of ensemble-
averaging process, further unknowns are introduced into the momentum equations by means of
Reynolds stresses ͞u´iu´j . In engineering flows, closure approximation using two-equation models
for u
͞ ´ju´j have gained popularity because of their simplicity. In this article the study is confined to the
κ-ε model, which employs additional transport equations for turbulent kinetic energy κ and its
dissipation rate ε, and they are given as

κ κu j u   ν t  κ 
  ui uj i  ε    ν  
t x j x j x j  σ κ  x j 
Eq. 4.4
ε εu j ε ε2   ν t  ε 
  
 C1 u i u j  C 2    ν  
t x j κ κ x j  σ ε  x j 

Where κ is the turbulent kinetic energy, is the turbulent kinetic energy dissipation rate, ν is the fluid
kinematic viscosity, and νt is the eddy viscosity. C1, C2, σκ, and σε are the model constants given by
Table 3.1.

4.2.1.3 Numerical Strategies


The governing equations for velocities and turbulent quantities are solved using the finite-volume-
based commercial solver FLUENT® 6.2. The equations are discretized on a collocated grid in fully
implicit form. Momentum equations are solved using the QUICK scheme, and the SIMPLE algorithm
is used for coupling the pressure and velocity terms. The second-order upwind scheme is used to
discretize convective terms and also the terms in equations for turbulent quantities. The second-
order implicit scheme is used for time integration of each equation. The present nonlinear model is
incorporated in FLUENT through User-Defined Functions (UDFs). The nonlinear stress term is added
as a source term in equations for κ and ε.
The turbulent viscosity is also made to vary. The implementation of the present Nonlinear κ-ε model
(hereafter referred to as NLKE) was validated for turbulent flow around a square cylinder. A box-
type computational domain (Figure 4.1) with structured grid in Cartesian coordinates having origin
at the center of the cylinder is used. Stream wise direction is along the x axis with x ¼ 0 at the center
of the cylinder, the y axis is the vertical axis with y ¼ 0 the wake centerline, and the z axis is the span
wise direction with z ¼ 0 being the mid span of the cylinder. The upstream boundary with uniform
inlet velocity is placed at 7D from the center of the body, where D is the diameter of the cylinder. This
extent is less than that used in the LES simulation of FF02, where the inlet boundary was kept at 10D
from center of the body and the URANS, and the LES simulation of LU01, who used 8D from the center
of the body. At the inlet boundary, 2% turbulence intensity is specified. The convective boundary
condition at the outlet is specified at 20D downstream of the body. The same domain size was used
by FF02, whereas LU01 used 24D. In the y direction, slip boundary condition is applied at 8D from
the center of the body, which is the same as the one used by previous references. Most other reported
LES simulations have used either an O-grid or a C-grid with larger boundary dimensions. All the LES
simulations were done with span wise domain size of PD, whereas the URANS simulations have been
reported only for two-dimensional flow. The three-dimensional stability analysis and the experiment
on a subcritical Reynolds number showed the dominant span wise scales having wavelengths of
approximately three to four cylinder diameters in the Reynolds number range 180 < Re < 240. After
this Reynolds number, the wavelength shortens to nearly one diameter. In our simulation, we have
taken a span wise length of 4D along the z direction, and periodic boundary condition is enforced on
the boundary. This extent is slightly larger than that used in LES and DNS calculations. A non-
equilibrium wall function approach is used to capture the adverse pressure gradient effect.
49

4.2.1.4 Geometry and Meshing


The present NLKE model was simulated with five different grids for grid dependence studies. In all
grids, along the circumference of the cylinder, 144 mesh points were placed, and in the cross-stream
direction, 85 mesh points were used. For all grids, the mesh points were placed uniformly in the span
wise direction. Different grids were achieved by systematically changing the resolution in the stream
wise and span wise directions. Figure 4.1 shows the schematic representation of the computational
domain used. Due to ease of domain, a simple structured grid being used. Of course, to be noted that
the overall grid size is coarser than those used by LES and DNS. The same grid is used to simulate
three-dimensional flow by the
standard κ-ε model (referred
to as 3DSKE) as well as two-
dimensional flow by the
nonlinear κ-ε model (2DNLKE)
for comparison purposes.

4.2.1.5 Results and


Discussion
The solution is initiated and
allowed to march in time with
a non-dimensional increment
of dt (Δt U∞/D =5x103) until
the vortex shedding becomes
periodic. The time variation of
the lift coefficient is shown in
Figure 4.2, and periodicity is
observed. The simulation is
continued for five more vortex
shedding cycles to advect all
the numerical errors to
downstream. In Breuer’s LES Figure 4.1 Schematic Representation of the Computational Domain
work 69, in the time history of
the lift coefficient, there was a low-frequency component over a regular periodic component. Hence,
in order to achieve reproducible statistics, averaging was done over 22 vortex shedding cycles. In the
present work, as there are no such features, the time averaging is done over 10 vortex shedding cycles
to obtain both bulk and mean field
quantities. Here T (t U∞/D) is non
dimensional time for one vortex shedding
cycle.
The present three-dimensional simulation
with the NLKE model (referred as 3DNLKE)
shows better agreement with the
experimental results. The present 3DNLKE
simulation predicted the mean drag
coefficient (Cd mean) and Strouhal number
(St) within the experimental uncertainty, but
the two-dimensional simulation (2DNLKE)
under predicts these values. The Figure 4.2 Variation of lift Coefficient with Time

69M. Breuer, “Large Eddy Simulation of the Subcritical Flow past a Circular Cylinder: Numerical and Modeling
Aspects”, Int. J. Numerical Meth. Fluids, vol. 28, pp. 1281–1302, 1998.
50

recirculation length (Lr= D) and negative velocity (U1 min= U∞) in the wake predicted by the present
3DNLKE simulation agreed well with those of experimental and LES results. Simulation by 3DSKE
predicts longer recirculation length
(Lr=D) and less negative velocity in the
wake. Location of the minimum velocity
(rmin=D) by the present two-dimensional
simulation with the NLKE model
(2DNLKE) and the two-dimensional
URANS simulation result from LU01
showed an under predicting trend.
However, the present 3DNLKE results
match well with experiment. The mean
stream wise velocity recovery along the
wake centerline is shown in Figure 4.4.
It can also be observed that the
experimental data showed some scatter.
Except for these locations, the present
3DNLKE results are in good agreement
with experiment. However, slower wake Figure 4.3 Comparison of Wall Pressure Coefficient (Cp)
velocity recovery is observed with 3DSKE
model. at this Reynolds number and
beyond ineffective, as variations in
the span wise direction are neglected.
Figure 4.3 provides a comparison of
the wall pressure coefficient with
experimental and different numerical
simulations. The LES results agree
well with the experimental results,
but all our three simulations (3DSKE,
2DNLKE, and 3DNLKE) show
different trends. The over prediction
of stagnation-point value by the SKE
model is a well-known fact and is
attributed to the steep velocity
gradient on the upstream side. The Figure 4.4 Streamline Patterns at Four Different Time
2DNLKE and 3DNLKE models predict Intervals for one Vortex Shedding Cycle
the stagnation point correctly, but the
maximum negative pressure and base pressure coefficient differ from the experimental results
predicts the trend.

4.2.1.6 Conclusions
In the present work, 3-D unsteady computation of flow past a circular cylinder at subcritical Reynolds
number has been performed using a nonlinear κ-ε model to evaluate its applicability. The same test
case was simulated with 2DNLKE and its 3D counterpart 3DSKE to understand the effectiveness of
the present model. The bulk parameters and the wake velocity recovery match well with
experimental data and LES results. Since the grid requirement is not as severe as in LES and the
number of cycles required to do averaging is also less, computational cost associated with the present
model is very much less. For high-Re flows and flows encountered in practical engineering
applications, there is a restriction on the mesh size and the LES technique is prohibitively expensive.
Encouraging performance of the present NLKE model suggests that it could be used as an alternative
51

tool in these situations. Further improvement in the prediction may be possible by making the model
fully cubic form.

4.3 2nd Order Closures (Reynolds Stress Models - RSM)


Due to the increased complexity of this class of turbulence models, 2nd order closure models do not
share the same wide use as the more popular two-equation or algebraic models. The most
noteworthy efforts in the development of this class of models was performed by [Donaldson and
Rosenbaum (1968), Daly and Harlow (1970), and Launder , Reece, and Rodi (1975)]. The latter has
become the baseline 2nd order closure model, with more recent contributions made by [Lumley
(1978)], [Speziale (1985, 1987), and many other thereafter, who have added mathematical rigor to
the model formulation. The first three belong to so called 1st order closures. LRR R-ε and Launder-
Gibson R-ε models, calculates the anisotropic Reynolds stresses present in typical flows of complex
strain fields or significant body forces. RSM is the most complex of the turbulence models, and was
designed to address the problems of the κ-ε model (which cannot predict flows in long non-circular
duct because of the isotropic modelling of the Reynold’s stresses). The RSM can therefore accurately
account for the Reynold’s stress field directional effects. Because of the many Reynold’s stresses to
model, there are seven extra partial differential equations to solve, making computing costs very
high. The Reynolds Stress models, sometimes called stress-equation models, are enforcing to those
models which do not assume the Boussiqes assumption. Thus it would seems that these are perhaps
the ultimate models. Nevertheless these models still utilize approximation and assumption in the
modeling terms.

4.4 LES Model


The principal idea behind LES is to reduce the computational cost by ignoring the smallest length
scales, which are the most computationally expensive to resolve, via low-pass filtering of the Navier–
Stokes equations. Such a low-pass
filtering, which can be viewed as a time-
and spatial-averaging, effectively removes
small-scale information from the
numerical solution. This information is
not irrelevant, however, and its effect on
the flow field must be modeled. A task
which is an active area of research for
problems in which small-scales can play
an important role, such as near-wall
flows , reacting flows, and multiphase
flows70. Large-scale motion is calculated in
LES, while the small-scale motion needs to Figure 4.5 Difference Between the Filtered Velocity
be modeled because of the effects of large- and the Instantaneous Velocity
scale motion. The most important aspect
in application of LES is the use of suitable Sub-Grid Scale model. The implication to [Kolmogorov's
(1941)] theory of self-similarity is that the large eddy of simulation are dependent on geometry,
while the smaller scales more universal. This feature allows one to explicitly solve for the large eddies
in a calculation and implicitly account for the small eddies by using a Sub-Grid-Scale (SGS) model.
LES always solves three-dimensional, time dependent flow, calculating a mean of time-dependent
flow fields. Therefore, it is best suited for transient simulations. The main difference between

70 Wikipedia.
52

conventional turbulence modeling and LES is the averaging procedure. The LES technique does not
involve the use of ensemble average; rather it consists in applying a spatial filter to N-S equations.

Filter Definition
Mathematically, one may think of separating the velocity field into a resolved and sub-grid part. The
resolved part of the field represent the "large" eddies, while the sub-grid part of the velocity
represent the "small scales" whose effect on the resolved field is included through the sub-grid-scale
model71. This is called explicit filtering and Figure 4.5 illustrates the difference between the filtered
velocity ūi and the instantaneous velocity ux. formally, one may think of filtering as the convolution
of a function with a filtering kernel G:

ui (x⃗) = ∫ G(x⃗ − ξ)u(ξ)d(ξ) ui = u̅i + u′i


Eq. 4.5
Where ūi the resolvable scale, and u’i is the sub-grid scale part. However, most practical (and
commercial) implementations of LES use the grid itself as the filter (the box filter) and perform no
explicit filtering. This is best shown in series of Error! Reference source not found. where L defines
he largest eddies and Δ is the box filter based on mesh size. It is important to note that the large eddy
simulation filtering operation does not satisfy the properties of a Reynolds operator. The expression
most often used for
sub-grid scale part as a geometric mean is

g  ( x  y  z )1/3 Eq. 4.6

or its generalization, the cubic root of the cell volume. In case of anisotropic grids, definition tends to
provide a fairly low value72. For this reason, the quadratic mean is used as following

1/2
 2x  2y  2z 
g    Eq. 4.7
 3 
 
which is advocated in some publications. Other authors favor the maximum

 g  Max ( x ,  y ,  z ) Eq. 4.8

Filtered Incompressible N-S Equations


Using Einstein notation, the Navier–Stokes equations for an incompressible fluid in Cartesian
coordinates are:

u i u i u i u j 1 p  2ui
0 ,   ν Eq. 4.9
x i x j x j ρ x i x jx j

71From Wikipedia.
72Jochen Frӧhlich , Dominic von Terzi,” Hybrid LES/RANS methods for the simulation of turbulent flows”,
Progress in Aerospace Sciences 44 (2008) 349– 377.
53

Filtering the Momentum equation results in

u i u i u j 1 p  2u i
  ν or
t x j ρ x i x jx j
Eq. 4.10
 u i u i u j 1 p 2 ui
  ν
t x j ρ x i x jx j

This equation models the change of time of the filtered variable u͞i . Since the unfiltered variable ui
are not known, it is impossible to directly calculate:

∂u
̅̅̅̅̅
i uj ∂u̅i u̅j

⏟∂xj ⏟∂xj
Not Known Known
Eq. 4.11
However the quantity on the right is known. Substituting:

∂u̅i ∂u̅i u̅j 1 ∂p̅i ∂2 u̅i ∂u


̅̅̅̅̅
i uj ∂u̅i u̅j
+ =− +ν − ( − )
∂t ∂xj ρ ∂xi ∂xj ∂xj ⏟ ∂xj ∂xj
∂τij
∂xj
where τij = u
̅̅̅̅̅
⏟ i uj − u
̅ i u̅j
SGS Modeling
Eq. 4.12
The result is a set of LES equations, resulting in one SGS turbulence model which can be formulated
as the following73:

1 1 ∂u̅i ∂u̅j
τij = −2μSGS S̅ij + τij δij where S̅ij = ( − )
3 2 ∂xj ∂xi
Eq. 4.13
Where SGS is the artificial or the sub-grid scale viscosity which acts as the constant of proportionality
and Śij is the average strain rate. The size of the SGS eddies are determined by the filter choice as well
as the filter cut-off width which is used during the averaging operation. The SGS viscosity can be
obtained by the following semi-empirical formulation:

μSGS = ρ(CSGS ∆g )2 |√2S̅ij S̅ij |


Eq. 4.14
where Δg as described before is the filter cut-off width, and the CSGS empirical constant which is
usually specified in a range, 0.19 < CSGS < 0.24.74

73 T. Ganesan and M. Awang, ”Large Eddy Simulation (LES) for Steady-State Turbulent Flow Prediction”, Springer
International Publishing Switzerland 2015.
74 See Previuos.
54

Numerical Methods for LES


Large eddy simulation involves the solution to the discrete filtered governing equations using CFD.
LES resolves scales from the domain size down to the filter size Δ and as such a substantial portion
of high wave number turbulent fluctuations must be resolved. This requires either high-order
numerical schemes, or fine grid resolution if low-order numerical schemes are used. [Pope]75
addresses the question of how fine a grid resolution is needed to resolve a filtered velocity field u(x).
[Ghosal]76 found that for low-order discretization schemes, such as those used in finite volume
methods, the truncation error can be the same order as the sub-filter scale contributions, unless the
filter widths considerably larger than the grid spacing. While even-order schemes have truncation
error, they are non-dissipative,77 and because sub-filter scale models are dissipative, even-order
schemes will not affect the sub-filter scale model contributions as strongly as dissipative schemes.

Implicit vs Explicit Filtering


The filtering operation in large eddy simulation can be implicit or explicit. Implicit filtering
recognizes that the sub-filter scale model will
dissipate in the same manner as many numerical
schemes. In this way, the grid, or the numerical
discretization scheme, can be assumed to be the LES
low-pass filter. While this takes full advantage of the
grid resolution, and eliminates the computational cost
of calculating a sub-filter scale model term, it is
difficult to determine the shape of the LES filter that is
associated with some numerical issues. Additionally,
truncation error can also become an issue78. In
explicit filtering, an LES filter is applied to the
discretized Navier–Stokes equations, providing a
well-defined filter shape and reducing the truncation
error. However, explicit filtering requires a finer grid
than implicit filtering, and the computational cost
increases with [Sagaut]79 covers LES numeric in
greater detail. [Sarkar and Voke]80 carried out an LES
study of interactions of passing wakes and in flexional
boundary layer over a low-pressure turbine blade. Figure 4.6 Vorticity Prompted by the Wake
Figure 4.6 shows flow structures due to the complex Passing Cycle
interactions of passing wakes and the separated shear
layer where the iso-surface of vorticity at an instant of time through the wake passing cycle.

Quantitative Aspects of Comparison of RANS vs. LES Models


There are numerus studies done on comparison of advanced RANS models against large eddy

75 Pope, S. B. (2000). Turbulent Flows. Cambridge University Press.


76 Ghosal, S. (April 1996). "An analysis of numerical errors in large-eddy simulations of turbulence". Journal of
Computational Physics 125 (1): 187–206.
77 Randall J. Leveque (1992), “Numerical Methods for Conservation Laws (2nd Ed.)”, Birkhäuser Basel. ISBN 978-

3-7643-2723-1.
78 Grinstein, Fernando, Margolin, Len, Rider, William, “Implicit large eddy simulation”, Cambridge University

Press. ISBN 978-0-521-86982-9, 2007.


79 Sagaut, Pierre, “ Large Eddy Simulation for Incompressible Flows”, (3rd Ed.), Springer, 2006.
80 “Large-eddy simulation of unsteady surface pressure over a LP turbine due to interactions of passing wakes and

in flexional boundary layer” , Journal of Turbomachine, pp. 221–23, 2006.


55

simulation (LES), chief among them is [Keshmiri et al.]81 etc. The biggest difference between LES and
RANS is that, contrary to LES, RANS assumes that ūi = 0 (see the Reynolds-averaged Navier–Stokes
equations). In LES the filter is spatially based and acts to reduce the amplitude of the scales of motion,
whereas in RANS the time filter removes ALL scales of motion with timescales less than the filter
width. What distinguishes LES from RANS is the definition of the small scales. While LES assumes the
small scales to be smaller than the mesh size Δ, RANS assumes them to be smaller than the largest eddy
scale L. So the quality of LES model is directly dependent in mesh size Δx. The mathematical similarity
of LES and RANS equations, as evidence in the equations (4.15), are being solved essentially the
same. However, the physics are different.

RANS
u
t
 
   ( u u )    ν  ν T   u   u
T
 
 p 
LES
u
t

   (u u )    ν  ν SGS   u   u T
 
 p 
Eq. 4.15
It is obvious, the only change is in the dynamic viscosity determination, νT and νSGS. Or, the main
difference being that in RANS the unclosed term is a function of the turbulent kinetic energy and the
turbulent dissipation rate whereas in
LES the closure term is dependent on the
length scale of the numerical grid. So in
RANS
RANS the results are independent of
the grid resolution! and usually the DES
needs more refine mesh that RANS.
Another point of view is that RANS can
only give a time averaged mean value
for velocity field since it is based on DES
time averaging. In fact velocity field in
this method is averaged over a time
period of "Δt" which is considerably
higher than time constant of velocity Figure 4.7 Time Averaged (RANS) vs Instantaneous (DES)
fluctuations. An example would be the Simulation Over a Backup Step
flow in backward step using both RANS
and DES models, (see Figure 4.7) where the difference is obvious. While DNS resolves all scales of
motion, all the way down to the Kolmogorov scale, LES is next up and resolves most of the scales,
with the smallest eddies being modeled. RANS is on the other end of the spectrum from DNS, where
only the large-scale eddies are resolved and the remaining scales are modeled.

Motivations for Coupling Methods Between RANS and LES


As indicated previously, (RANS) models provide results for mean quantities with engineering
accuracy at moderate cost for a wide range of flows. In other situations, dominated by large-scale
anisotropic vortical structures like wakes of bluff bodies, the average quantities are often less
satisfactory when a RANS model is employed. Then large eddy simulation (LES) performs generally
better and bears less modeling uncertainties. Furthermore, LES by construction provides unsteady
data that are indispensable in many cases: determination of unsteady forces, fluid–structure
coupling, identification of aerodynamic sources of sound, and phase-resolved multiphase flow, to

81Amir Keshmiri, Osman Karim, Sofiane Benhamadouche, “Comparison of Advanced Rans Models Against Large
Eddy Simulation And Experimental Data In Investigation Of Ribbed Passages With Heat Transfer”, The 15th
International Conference on Fluid Flow Technologies, Budapest, Hungary, September 4-7, 2012.
56

name but a few issues [Frohlich & Terzi]82. Unfortunately, LES is by a factor of 10 to 100 more costly
than RANS computations; LES requires a finer grid, cannot benefit from symmetries of the flow in
space, and provides mean values only by averaging the unsteady flow field computed with small time
step over a long sampling time. Hence, it seems natural to attempt a combination of both turbulence
modeling approaches and to perform LES only where it is needed while using RANS in regions where
it is reliable and efficient.
Another and somewhat different motivation for LES/RANS coupling stems from wall-bounded flows.
Close to walls, the LES philosophy of resolving the locally most energetic vortical structures requires
to substantially reduce the step size of the grid since the dominating structures become very small in
this region. Furthermore, when increasing the Reynolds number, the scaling of the computational
effort is similar to that of a DNS in its dependence on Re just with a smaller constant. That makes the
approach unfeasible for wall-bounded flows at high Re, such as the flow over a wing . As a remedy,
some sort of wall model can be introduced to bridge the near-wall part of the boundary layer and to
make the scaling of the required number of grid points independent of Re. A RANS model depends
on physical quantities describing the entirety of the turbulent fluctuations. For the sequel it is
necessary to define the specifics of LES models and RANS models. Using an unsteady definition of a
Reynolds average as discussed above, the transport equations for the Reynolds-averaged velocity
<ui> read:

 ui  ui u j p   ν u i  τ ijRANS
   
t x j x j x j  x j 
 x j
Eq. 4.16
u i u i u j p   νu i  τ ij
DES

   
t x j x j x j  x j  x j

For example, the κ–ε model determines

  ui   u i 
τ ijRANS  f  , κ, ε, C  , τ ijDES  f  , Δ g , C  Eq. 4.17
 x i   x i 

where C is a model constant, κ the turbulent kinetic energy, and ε the turbulent dissipation rate. The
latter two are determined from other relations. For LES based on the (Smagorinsky) model uses a
relation like where Δg is a length scale related to the numerical grid, Since there exist many variants
of LES and RANS models we define the following: a model qualifies as an LES model if it explicitly
involves in one or the other way the step size of the computational grid. RANS models, in contrast,
only depend on physical quantities, including geometric features like the wall distance.

4.4.6.1 Principal Approaches to Coupling LES with RANS


The similarity of the equations and the considerations suggest the concept of unified modeling. This
approach is based on using the same transport equation for some resolved velocity ūi, yet to be
specified in its meaning:

82Jochen Frӧhlich , Dominic von Terzi,” Hybrid LES/RANS methods for the simulation of turbulent flows”,
Progress in Aerospace Sciences 44 (2008) 349– 377.
57

u i u i u j p   νu i  τ ij
Model

    Eq. 4.18
t x j x j x j  x j  x j

A transition from LES to RANS can be achieved in several ways. One possibility is blending, i.e. by a
weighted sum of a RANS model and an LES the models according to

 ijModel  f RANS ijRANS  f LES ijLES Eq. 4.19

In this fashion, f RANS and f LES are local blending coefficients determined by the local value of a given
criterion. According to [Faridul Alam]83, a linear blending of f would be sufficient with the
proportionality constants can be both spatially and temporally varying, Which bring the grid
sensitivity issue to the picture. In that case, the residual stress term can be expressed as a weighted
average of both the SGS and RANS stress as follows. Another strategy is to use a pure LES model in
one part of the domain and a pure RANS model in the remainder, so that a boundary between a RANS
zone and an LES zone can be specified at each instant in time. The transport equation for the velocity,
however, is the same in both zones with no particular adjustment other than switching the model
term at the interface. This way the
computed resolved velocity is
continuous. We term this strategy
Interfacing LES and RANS.
Furthermore, if the interface is
constant in time, it is called a hard
interface. If it changes in time
depending on the computed solution, it
is termed a soft interface.
Segregated modeling is the counterpart
to unified modeling as LES is employed
in one part of the computational
domain, while RANS is used in the
remainder. With segregated modeling,
however, the resolved quantities are
no more continuous at the interfaces. Figure 4.8 Integrated RANS-LES Computations in Gas
Instead, almost stand-alone LES and Turbines: Compressor-Diffuser,
RANS computations are performed in
their respective subdomains which are then coupled via appropriate boundary conditions. Except for
laminar flows, the solution is discontinuous at these interfaces. This avoids any gradual transition in
some gray area characteristic of unified turbulence models. Segregated modeling allows for
embedded LES by designing a configuration where in an otherwise RANS simulation a specific region
is selected to be treated with LES with full two-way coupling between the zones84.

4.4.6.2 Results and Discussion for Segregated Modeling


As an example of segregated modeling, is integrated RANS-LES of the NASA stage 35 compressor and
a diffuser. While the compressor stage is computed with the RANS flow solver, the subsequent

83 Mohammad Faridul Alam, David Thompson and Dibbon Keith Walters, “Critical Assessment of Hybrid RANS-
LES Modeling for Attached and Separated Flows”, IntechOpen, 2017.
84 Jochen Frӧhlich , Dominic von Terzi,” Hybrid LES/RANS methods for the simulation of turbulent flows”,

Progress in Aerospace Sciences 44 (2008) 349– 377.


58

diffuser with the LES flow solver (see Figure 4.8). Here, we look at the vorticity magnitude
distribution at the 50% clip plane of the stator. Again, we can identify the wakes of the stator passing
the interface. The different description of turbulence in the two mathematical approaches is
apparent. While on the RANS side the turbulence is modeled in a turbulence model and cannot be
seen in the vorticity distribution, on the LES side the fine scale turbulence is regenerated and can be
identified as small-scale structures in the LES solution85. Another example in the feasibility of a
hybrid RANS–LES approach is the numerical simulation of aircraft wing-tip vortices which has been
studied by [Kolomenskiy et al]86. Mesh sensitivity tests of our RANS solver and comparisons between
two different turbulence models indicate that the RANS approach adequately describes the flow
upstream from the trailing edge, but overestimates the rate of decay of the wing-tip vortex. A hybrid
RANS–LES method is presented that results in a better agreement with the wind tunnel experiment
for numerical simulation of the wake of an airliner.

4.4.6.3 Coupling in RANS/LES Interface


A fully coupled solution requires that all flow variables must be exchanged at the interface. When
some engine components are computed with LES and others with RANS approximations will have to
be made to couple instantaneous and averaged variables. To simplify the problem, we consider only
the one-way coupling of the velocity and turbulence variables. One-way coupling means that
information is passed only downstream; the variables at the inlet of the downstream domain are
computed from the variables at the outlet of the upstream domain. For the RANS/LES interface,
turbulent fluctuations need to be added to the velocity from an upstream RANS computation. For the
LES/RANS interface, such as the combustor/turbine interface, a simple time average of the velocity
provides a mean velocity at the inlet
of the RANS domain. This velocity
distribution is again highly non-
uniform, which allows to describe
turbulence at the inlet with the local
turbulence generation from the mean
velocity.
The flow in the compressor is
computed with unsteady RANS using
the k-ω model and the flow in the
combustor is computed with LES. The
last blade row of the compressor is a
row of stators and does not rotate.
The blade row upstream is a row of
rotors that rotate counter-clockwise.
The mean flow is highly complex, Figure 4.9 Compressor and combustor: RANS and LES axial
velocity, mid-passage (Courtesy of Medic et al.)
with the wakes originating from the
last stage of the compressor (see
Figure 4.9). These wakes are unsteady due to the rotation of the rotors in the compressor. Larger
values of turbulent kinetic energy are in the regions with strong velocity gradients. The large values
of k near the hub might be spurious; they are highly dependent on the quality of the grid. This
illustrates the fact k from the k-ω model usually fails to accurately represent the true turbulent
kinetic energy in complex flows. The flow field in the downstream LES domain is highly dependent

85 J. U. Schlüter, X. Wu, S. Kim, J. J. Alonso, and H. Pitsch, AIAA-2004-369, 42nd Aerospace Sciences Meeting and
Exhibit Conference, January 2004.
86 Dmitry Kolomenskiy, Roberto Paoli, and Jean-Franc¸ois Boussuge, “Hybrid Rans–Les Simulation of Wingtip

Vortex Dynamics”, FEDSM2014-21349.


59

on the conditions at the inlet. To generate an inflow profile for the LES in the combustor, the mean
velocity at the combustor inlet is set equal to the RANS velocity at the compressor exit and
appropriate fluctuations need to be added [Medic, et al]87.
The aim of segregated modeling is to compute all models in their regime of validity: steady RANS for
flows with stationary statistics and unsteady LES with high resolution where it is needed. Therefore
one can choose the best suited method for each subdomain without considering their compatibility
and without fear of inconsistencies in their use. Furthermore, any gray zone where the model is left
alone with generating fluctuations in some transition process is avoided. The price to pay is the need
for comparatively complex coupling conditions. For block-structured solvers, however, the routines
for data exchange required anyway facilitate a straightforward implementation. Inappropriate
coupling conditions lead to contamination of the results in the LES or RANS subdomains. Depending
on the type of the interface, the requirements on the coupling conditions (inlet and outlet) differs
[Frohlich & Terzi]88

4.5 DES Model


An alternative between LES and RANS models would the Detached Eddy Simulations (DES) which
attempts to combine the best
aspects of RANS and LES 13 M cells
methodologies in a single
solution strategy. This is more
pronounce in near-wall regions
in a RANS-like manner, and treat
the rest of the flow in an LES-like
manner. The model was
originally formulated by
replacing the distance function d
in the Spalart-Allmaras (SA)
model with a modified distance
function D as:

D  Min (d, CDES , ) 16 M cells


Eq. 4.20
Where CDES is a constant and Δ is
the largest dimension of the grid
cell in question. This modified
distance function causes the
model to behave as a RANS model
in regions close to walls, and in a
Smagorinsky-like manner (LES)
away from the walls. This is
usually justified with arguments
that the scale-dependence of the
model is made local rather than
global, and that dimensional Figure 4.10 DES and Effect of Grid Density on the Wake Flow

87 G. Medic, D. You AND G. Kalitzin, “An approach for coupling RANS and LES in integrated computations of jet
engines”, Center for Turbulence Research, Annual Research Briefs, 2006.
88 Jochen Frӧhlich , Dominic von Terzi,” Hybrid LES/RANS methods for the simulation of turbulent flows”,

Progress in Aerospace Sciences 44 (2008) 349– 377.


60

analysis backs up this claim. In summary, an example of a hybrid technique, detached eddy
simulation (DES) is a modification of a RANS model in which the model switches to a sub-grid scale
formulation in regions fine enough for LES calculations. Therefore, the grid resolution is not as
demanding as pure LES, there by considerably cutting down the cost of the computation. A study in
CD Adapco® shows that for wake flow drag calculation, DES is more accurate than RANS and highly
dependent in grid resolution (see
Figure 4.10, top 13 M cells vs bottom-16 M cells). Another example would be the presentation of
flow in rectangular ogive fore-body (Aircraft fore body)89. In that environment we set close to the
wall D is by the wall parallel spacing’s, i.e., D = d; d = << Δ for RANS and away the wall CDESΔ < D, i.e.,
D = CDES Δ for LES. Using an unstructured generated using VGRID. Simulation details can be found in
[Viswanathan, Squires and Forsythe 2006].

 Re = 2.21 x 106, Mach number = 0.21


 Grid sizes from 2.1 x 106 cells to 8.75 x 106 cells
 Angle of attack: 60 degrees

4.6 DNS Model


The DNS is a research tool, and not a brute-force solution to the Navier-Stokes equations for
engineering problems. It is a method which is CPU intensive, as it tries to resolve N-S equations
without any turbulence approximation, requiring a very fine numerical resolution to capture all the
details of turbulence. The theory is that the whole range of spatial and temporal scales of the
turbulence must be resolved. All the spatial scales of the turbulence must be resolved in the
computational mesh, from the smallest dissipative scales (Kolmogorov microscales), up to the
integral scale L, associated with the motions containing most of the kinetic energy. It is not hard to
prove that for turbulent Reynolds number, we have:

u L
N3DNS  Re9/4 Re 

Eq. 4.21
Where N is the number of points along a given mesh direction with increments h and ú is the root
mean square (RMS) of velocity. Hence, the memory storage requirement in a DNS grows very fast
with the Reynolds number90. In addition, given the very large memory necessary, the integration of
the solution in time must be done by an explicit method. This means that in order to be accurate, the
integration, for most discretization methods, must be done with a time step, Δt, small enough such
that a fluid particle moves only a fraction of the mesh spacing h in each step. Therefore, it remains
limited to very simple cases. Filtering requires a finer grid than implicit filtering, and the
computational cost increases with [Sagaut (2006)] covers LES numeric in greater detail91. [Sarkar
and Voke]92 carried out an LES study of interactions of passing wakes and in flexional boundary layer
over a low-pressure turbine blade and Figure 4.6 shows flow structures due to the complex
interactions of passing wakes and the separated shear layer. Evidently, as it appears in Eq. 4.21 and
above discussion, the DNS method is highly dependent to Reynolds number which will be debated in
the next section.

89 Kyle D. Squires, Les Application in Aerodynamics”, School of Mechanical, Aerospace, Chemical and Materials
Engineering Arizona State University Tempe, Arizona, USA.
90 From Wikipedia.
91 Sagaut, Pierre (2006). “Large Eddy Simulation for Incompressible Flows”, (3rd ed.), Springer.
92 “Large-eddy simulation of unsteady surface pressure over a LP turbine due to interactions of passing wakes and

in flexional boundary layer”, J Turbo machine, 128 (2) (2006), pp. 221–231
61

The Reynolds Number Constraint


Turbulence contains eddies with a wide range of sizes. These eddies interact with each other in a
non-linear fashion through their induced velocity fields, changing the orientation and shape of their
neighbors. As first described by Richardson (1922) and quantified by Kolmogorov (1941), the net
effect of this change-of-shape (i.e. straining) process is to ‘cascade’ kinetic energy from the largest to
the smallest scales of the turbulence. As a result, the largest eddies are the most energetic, and their
size, shape and speed are set by the details of the flow configuration, and not directly affected by the
viscosity of the fluid. The size of the smallest eddies, on the other hand, is determined both by how
much energy enters the cascade at the large scales and by the viscosity. The primary role of viscosity
is to define the scale at which the energy is dissipated. Therefore, the Reynolds number of the flow is
determines how small the smallest scales are, relative to the largest eddies93. This behavior, known
as Reynolds-number similarity, can be observed in Figure 4.11, which presents DNS results from the
same boundary-layer flow at two Reynolds numbers that differ by a factor of two. This illustrates the
challenge faced by DNS, which must use a domain large enough to comfortably include the largest
naturally-occurring eddies while using a grid fine enough to fully resolve the dissipation scales. Using
today’s most capable computers, this can only be done for Reynolds numbers orders of magnitude
smaller than found, for example, in full-scale aeronautical flows.

Figure 4.11 Vorticity contours from spectral DNS at two Reynolds numbers (Spalart et al. 2008).

Numerical Considerations
The obligation of having to resolve all spatial and temporal scales of the turbulence requires that
numerical errors be monitored and controlled. As a result, DNS has historically not used commercial
CFD packages, but specially-written codes, optimized for the flow-types of interest94. The need for
DNS algorithms to be efficient; that is, to have a high ratio of accuracy to computational cost which is
particularly important. There are a number of strategies that DNS codes have employed to do this,
including finite-volume, finite-element, and discrete-vortex as well as B-spline methods. But central
to development of an efficient algorithm for DNS are two methods; Spectral and Finite Difference.

93 Gary N. Coleman and Richard D. Sandberg, “A Primer on Direct Numerical Simulation of Turbulence – Methods,
Procedures and Guidelines”, Technical Report AFM-09/01a - March 2010.
94 Gary N. Coleman and Richard D. Sandberg, “A Primer on Direct Numerical Simulation of Turbulence –

Methods, Procedures and Guidelines”, Technical Report AFM-09/01a (March 2010).


62

Figure 4.12 Sound Radiated by a Mach 1.9 Circular Jet

4.6.2.1 Spectral Methods


Spectral methods approximate the flow variables as linear combinations ‘expansions’ of global basis
functions that involve complex-exponential or orthogonal polynomial Eigen solutions of an
appropriate problem. (See Figure 4.11). An exciting new development has been the field of
computational aeroacoustics, where both the fluid motion and the sound it radiates are directly
computed (see Tam95, Lele96). Computational aeroacoustics is still in its infancy; the sound from a
canonical flow such as a perfectly expanded supersonic jet has only just been computed [Freund, Lele
& Moin] as a unpublished information97. Figure 4.12 shows results for a Mach 1.9 jet from their
massive computations using 640x270x128 mesh points. Contours of vorticity (black) are overlaid on
dilatation contours (gray). The computations were performed by [Freund, Lele & Moin].

4.6.2.2 Finite Differencing


Because of their ease of implementation, suitability to parallelization, and possible high-order
accuracy, finite-difference (FD) schemes for DNS have become increasingly popular, especially for
the emerging area of computational aeroacoustics (CAA) (see Figure 4.13 where Instantaneous
contours of stream wise density gradients from high-order finite-difference DNS of a supersonic
axisymmetric wake at Mach 2.46 and Re= 100, 000 (based on freestream velocity and diameter of
body; Sandberg 2008). A wide range of options is available. Low-order FD methods allow complex
geometries and irregular grids, and in this sense are similar to finite-volume methods. The
computational efficiency of low-order (especially upwind) FD approximations (FDA) is often
unacceptable, requiring many more grid points than a spectral method to achieve the same accuracy.

95 TamCKW. “Computational aeroacoustics: issues and methods”, AIAA pap. 95–0677.


96 Lele SK. 1992, “Compact finite difference schemes with spectral-like resolution”, J. Comp. Phys. 103:16–42.
97 Parviz Moin and Krishnan Mahesh, “ Direct Numerical Simulation: A Tool in Turbulence Research”, Annu. Rev.

Fluid Mech. 1998. 30:539–78. Copyrightc 1998 by Annual Reviews Inc.


63

Figure 4.13 Instantaneous Contours of Stream-Wise Density Gradients from DNS

4.7 Strategies for Turbulence Modelling


According to [Spalart]98 turbulence predictions in aerodynamics face two principal challenges: (I)
growth and separation of the boundary layer, and (II) momentum transfer after separation. (I) is
simpler, but makes very high accuracy demands, and appears to give models of higher complexity
little advantage. (II) is now the arena for complex RANS models and the newer strategies, by which
time-dependent three-dimensional simulations are the norm even over two-dimensional geometries.
In some strategies, grid refinement is aimed at numerical accuracy; in others it is aimed at richer
turbulence physics. In some approaches, the
empirical constants play a strong role even when
the grid is very fine; in others, their role
vanishes. For several decades, practical methods
will necessarily be RANS, possibly unsteady, or
RANS/LES hybrids, pure LES being unaffordable.
Their empirical content will remain substantial,
and the law of the wall will be particularly
resistant. Estimates are offered of the grid
resolution needed for the application of each
strategy to full-blown aerodynamic calculations,
feeding into rough estimates of its feasibility
date, based on computing-power growth. The
numerical strengths of CFD increase by the year
thanks to the progress of computers, whereas
turbulence modelling almost stagnate. A
primary purpose here is to provide a viewpoint
on these relatively new, evolving and Figure 4.14 Stream Lines in Square Channel
misunderstood methods, besides predicting that with Nonlinear Constitutive –( Courtesy of
they will proliferate and make a major [speziale, 1987])
contribution.

98P.R. Spalart, “Strategies for turbulence modelling and simulations”, International Journal of Heat and Fluid
Flow 21 (2000) 252±263.
64

Physical Aspects (Modification to Simple RANS Models)


The field of classical RANS turbulence modelling is active. We are referring here primarily to eddy-
viscosity models. Improvements will be made to the simpler transport models, typically by adding
new empirical terms aimed at compressibility, streamline curvature or better anisotropy of the
Reynolds-Stress tensor (non-Boussinesq constitutive relations). Which can, for instance, create
secondary flows of the second kind in a square pipe [Speziale]99. An example is given in Figure 4.14.
The SA model was used with the non-linear constitutive relation, as a first attempt to improve on the
eddy-viscosity approximation. The SA model exceeded expectations, and is now useful to a rather
large user base. The model is not sophisticated, but it is not inaccurate.

Complex RANS Models


The DES on two grids also depicts the inclusion of smaller eddies allowed by the finer grid. The flow
is at a Reynolds number of 50,000. It led us to propose Detached-Eddy Simulation (DES), a further
step in the hybridization of LES (Spalart et al., 1997). The idea is to entrust the whole boundary layer
(populated with ``attached'' eddies) to a RANS model, and only separated regions (“detached''
eddies) to LES. It is aimed primarily at external flows. solutions are under-resolved in the separated
regions. Figure 4.15 illustrates the difference in description for the flow past a circular cylinder. The
runs on the right are DES, but the visible eddies are in the LES region. Of course the DES figures show
only a plane out of the 3D domain. Each step from SRANS through 2D URANS to LES/DES adds a
dimension: first the time and then the third space dimension. The cost increase is of an order of
magnitude each time. In return, the SRANS drag is too low at Cd ≈ 0.9, the URANS drag is much too

Figure 4.15 Simulation of Flow Past Circular Cylinder by Various Approaches (Shur et al., 1996;
Travin et al., 2000)

99 Speziale, C.G.,” On nonlinear K- l and K - ԑ models of turbulence”. J. Fluid Mech. 178, 459, 1987.
65

high at Cd ≈1.7, and the DES drag is in much better agreement with experiment, although grid effects
are still solutions are under-resolved in the separated regions. present: Cd ≈ 1.05 on the coarse grid
but 1.32 on the fine grid [Travin et al.]100. The experiment gives 1.2. For a compete discussion, please
refer to [P.R. Spalart]101.

Role of Grid Refinement


In RANS, the equations possess a smooth exact solution, and the numerical solution approaches that
solution as we refine the grid. The aim of grid refinement is numerical accuracy. In contrast, in LES
as it is practiced and in DES, the Sub-Grid-Scale (SGS) model adjusts to the grid so that the smallest
resolved eddies match the grid spacing. In a finer grid, resolving eddies to a smaller size gives the
large energy-containing eddies more eddies for genuine nonlinear interactions, making them more
accurate. The aim of grid refinement is now physical instead of numerical, to use simple words. This
distinction is tracked in Error! Reference source not found. (several methods had to be labelled
hybrid”, as their aim is different in different flow regions). Another description of it is that when the
aim is numerical, the turbulence model does not depend on the grid spacing but when the aim is
physical, it does. A consequence is that in URANS, no amount of grid refinement will override the
influence of the empirical content of the turbulence model. In contrast, in a method with the
``physical'' aim, grid refinement weakens the role of the modelled eddies and thus improves the
fidelity of the simulation. It has been proposed not to automatically link the width of the LES filter
and the grid spacing, in order to obtain solutions of the filtered equations free of significant numerical
errors. Then, the numerical errors remain the same fraction of the SGS kinetic energy, and it is fair to
write that “numerical and SGS effects cannot be separated”. This situation is disturbing to some careful
people, who would prefer to understand the physical system of the filtered equations, and then obtain
very accurate solutions to it. Widening the filter raises the magnitude of the sub-grid is effect can be
offset by the reduction of numerical errors, or even by the design of a superior stresses, which are
notoriously inaccurate locally. It is unlikely that SGS model, possibly gained from the better physical
understanding. On a given grid, an LES with a wide filter would cost almost the same as one with a
narrow filter: the cost per step would be the same, and the time step could only rise by a modest
amount. Conclusive tests of the filter-grid relationship would crucially depend on the definition of a
figure of merit.

Outlook
Progress in numerical methods and computers is intensifying the challenge for turbulence
treatments, to provide a useful level of accuracy in slightly or massively separated flows over fairly
complex geometries at very high Reynolds numbers. This is desirable in the near future, especially
the jet-engine industry. In addition, the needs of non-specialist users and automatic optimizers
dictate a very high robustness. Flows with shallow or no separation appear to be within the reach of
the current steady RANS methods or their finely calibrated derivatives, incorporating modest
improvements such as nonlinear constitutive relations. For such flows, transition prediction with
generality, accuracy, and robustness may prove more challenging than turbulence prediction. With
massive separation, it appears possible we will give up RANS, steady or unsteady. This will probably
be the major debate of the next few years. The alternative is a derivative of LES, in which the largest,
unsteady, geometry-dependent eddies are simulated and (for most purposes) “discarded” by an
averaging process. We have to balance our ambitions with cost considerations, and a table brief the
issue was tentatively provided.
A major consideration is whether LES is practical for the entire boundary layer, and it was strongly

100 Travin, A., Shur, M., Strelets, M., Spalart, P.R. “Detached-eddy simulations past a circular cylinder”. Flow,
Turbulence Combustion, 2000.
101 P.R. Spalart, “Strategies for turbulence modelling and simulations”, International Journal of Heat and Fluid

Flow 21 (2000) 252±263.


66

argued that this will not be the case, in the foreseeable future. This forces hybrid methods, with quasi-
steady RANS in the boundary layer. In this paper, LES was effectively defined as a simulation in which
the turbulence model is tuned to the grid spacing, and RANS as the opposite. Other more subtle
definitions probably exist, but this one seems to classify almost all the studies to date. Speziale's
hybrid proposal involves the grid spacing and the Kolmogorov length scale but, surprisingly, not the
internal length scale of the RANS turbulence model; thus, it is difficult to classify [Speziale]102. The
proposal of [Aubrun et al.]103. is also hybrid, as it leads to combining “modelled” and “resolved”
Reynolds stresses, but the modelled stresses do not scale (and vanish) with the grid spacing as they
do in LES 104. Variations on the now-running DES proposal clearly have a wide window of
opportunity.
The plausible spread of hybrid methods highlights the permanence of a partnership between
empiricism and numerical power in turbulence prediction at full-size Reynolds numbers. This
demands a balance in funding and in publication space. Since hybrid methods offer flexibility when
setting the boundary between “RANS region” and “LES regions”, the more capable the RANS
component is, the lower the cost of the hybrid calculation will be. Therefore, the switch to LES in
some regions does not remove the incentive to further the RANS technology. This scene also raises
the issue of which core of experiments and DNS will be the foundation of the empirical component of
the system. As ever, we will need simple flows for calibration of the RANS sub-system, and more
complex flows for validation of the full CFD system.

4.8 Modeling To Choose From


To recap, there are in general three categories of turbulence models to choose from, They are:

1. Reynolds-Averaged Navier-Stokes (RANS): Choosing an appropriated models, the


governing equations are solved in ensemble-averaged form, including appropriate models
for the effect of turbulence.

2. Large Eddy Simulation (LES): Large turbulent structures in the flow are resolved by the
governing equations, while the effect of the sub-grid scales (SGS) are modelled. The scale
separation is obtained by applying a filter to the governing equations which also influences
the form of the SGS models.

 Detached Eddy Simulation (DES): Hybrid method that treats near-wall regions with
a RANS approach and the bulk flow with an LES approach, (CPU intensive).

3. Direct Numerical Simulation (DNS): Resolves all scales of turbulence by solving the Navier-
Stokes equations numerically without any turbulence modelling, (CPU intensive).

102 Speziale, C.G., “Turbulence modeling for time-dependent RANS and VLES”, a review. AIAA J. 36 (2), 173,
1998.
103 Aubrun, S., Kao, P.L., HaMinh, H., Boisson, H, “The semi deterministic approach as way to study coherent

structures. Case of a turbulent flow behind a backward-facing step”, Proceedings of the Fourth International
Symposium on Engineering Turbulence Modelling and Measurements, Corsica, Elsevier, Amsterdam, 1999.
104 See Previous.
67
68

5 Classification of Turbulence Models


In general, the approach for solving turbulent flow equations can roughly be divided into four classes
as Direct Numerical Simulation (DNS), Reynolds Average N-S equations (RANS), Large Eddy
Simulation (LES) and Detached Eddy Simulation (DES), as 105:

1. RANS-based turbulence Models


a. Linear Eddy Viscosity models
i. Algebraic Models
1. Cebeci-Smith model
2. Baldwin-Lomax model
3. Johnson-King model
4. A roughness-dependent model
ii. One equation models
1. Prandtl's one-equation model
2. Baldwin-Barth model
3. Spalart-Allmaras model
4. Rahman-Siikonen-Agarwal Model
iii. Two equation models
1. κ-ε models
a. Standard κ-ε model
b. Realizable κ-ε model
c. RNG κ-ε model
2. κ-ω models
a. Wilcox's κ-ω model
b. Wilcox's modified κ-ω model
c. SST κ-ω model
b. Nonlinear Eddy Viscosity models
i. Explicit nonlinear constitutive relation
1. Quadratic κ-ε
2. Cubic κ-ε
3. Explicit Algebraic Reynolds Stress Models (EARSM)
ii. v2-f models
1. v2-f Model
2. ζ-f Model
c. Reynolds Stress Model (RSM)
2. Large Eddy Simulation (LES)
a. Smagorinsky-Lilly Model
b. Dynamic sub-grid-scale Model
c. RNG-LES Model
d. Wall-adapting local eddy-viscosity (WALE) Model
e. Kinetic energy sub-grid-scale Model
f. Near-wall treatment for LES Models
g. Structural Modeling
3. Detached Eddy Simulation (DES)
4. Direct Numerical Simulation (DNS)

105 Wikipedia.
69

5.1 Comparisons
Tableof
5.1various Turbulence
Advantages Models
& Disadvantages of Different Turbulence Models
Turbulence
Models Advantages Disadvantages

When ∂U/∂y = 0 ⇒ μ T = 0. Lack of


Zero-equation
transport of turbulent scales.
models, e.g. Cost-effective model applicable for a
Estimation of the mixing length is
mixing-length limited number of flows.
difficult. Cannot be used as a general
Model
turbulence model.
The use of an algebraic equation for the
One-equation
Cost-effective model applicable for a length scale is too restrictive.
models, e.g.
limited number of flows. Transport of the length scale is not
k-algebraic Model
accounted for.
Complete models in the sense that velocity
Two-equation Limited to an eddy-viscosity
and length scales of turbulence are
models, k–ε assumption. Turbulent viscosity is
predicted with transport equations. Good
Group standard, assumed to be isotropic. Convection
results for many engineering applications.
RNG, Realizable, k– and diffusion of the shear stresses are
Especially good for trend analysis. Robust,
ω and Mentor SST neglected.
economical and easy to apply.
Not good for round jets and flows
The most widely used and validated involving significant curvature, swirl
Standard k–ε Model
model. (isentropic), sudden acceleration,
separation and low-Re regions.
Modification of the standard k–ε model
Not as stable as the standard k–ε
RNG k–ε Model gives improved simulations for swirling
model. Not suited for round jets.
flows and flow separation
Modification of the standard k–ε model
Realizable k–ε gives improved simulations for swirling Not as stable as the standard k–ε
Model flows and flow separation. Can also handle models
round jets.
Works well at low Re. Does not need wall
Needs fine mesh close to the wall, first
κ–ω Model functions. Works well with adverse
grid point at y+ < 5.
pressure gradients and separating flow.
Uses k–ε in the free stream and k–ω in the Needs fine mesh close to the wall. Over
wall-bounded region. Works well with predicts turbulence in regions with
adverse pressure gradients and separating large normal strain, e.g. stagnation
SST Model
flow. Many authors recommend that the regions and regions with strong
SST model should replace the k–ε model as acceleration, but is better than k–ε.
the first choice.
Applicable for complex flow where the Computationally expensive with 11
turbulent-viscosity models fail. Accounts transport equations. Several terms in
Reynolds Stress for anisotropy. Good performance for many the transport equations must be
Models, (RSM) complex flows, e.g. swirl, flow separation closed. Poor
and planar jets. Performance for some flows due to the
closures introduced in the model.
Applicable to complex flows. Gives High computational cost. Large
information about structures in turbulent amount of data that must be stored and
Large-eddy
flows, ,as well as lots of information that post-processed. Difficult to find proper
Simulation (LES)
cannot be obtained otherwise. time-resolved boundary conditions for
flow.
No turbulence models are introduced. Extreme computational cost for
Useful at low Re Numbers, especially for practical engineering flow Simulations.
DNS
gaseous flows. Useful to develop and Huge amount of data
validate turbulence models
70

In the previous sections, we have presented turbulence models that are commonly encountered in
commercial CFD codes. In these sections the physical and mathematical principles underlying the
turbulence models were presented together with discussions about their limitations. An overview of
turbulence models, sorted by descending complexity, depicted on Table 5.1 which gives a short
summary of the advantages and shortcomings of these models106. In general the turbulence models
are developed to predict velocities accurately. The parameters in the models, (e.g., k and ε), may very
well be off by a factor of 3. Using these parameters in other models, e.g. for mixing or bubble break-
up, should be done with the awareness that the parameters do not have exact physical relevance but
only show the trend.

5.2 Numerical Considerations


Modern turbulence model equations pose special numerical difficulties that must be understood in
order to obtain reliable numerical solutions, even for boundary-layer flows where the equations are
parabolic. For one-equation, two-equation and second-order closure models, these difficulties
include stiffness caused by the presence of an additional time scale, singular behavior near solid
boundaries, and non-analytical behavior at sharp turbulent/non-turbulent interfaces . This chapter
focuses on these difficulties and on the solution methods for turbulence-model equations that have
evolved.

Elementary Time-Marching Methods


The examples of time marching methods and their effect on solution will not be discussed here since
it is already detailed in average flow analysis. Just a mentioned method will be suffice here.

 Explicit Schemes such as DuFort-Frankel, Godunov, Lax-Wendroff and Mac-Cormack


methods.
 Implicit methods such as Crank-Nicolson, Euler, and Alternating Direction Implicit (ADI)
schemes.

5.2.1.1 Block-Implicit Methods


The most efficient numerical methods currently available for complex flow fields are block-implicit
methods. They differ from elementary implicit methods in one very important respect. Specifically,
when an elementary implicit scheme is applied to a coupled set of equations, each equation is solved
in sequence. In the context of a system of equations, this is usually referred to as a sequentially-
implicit method. By contrast, a block-implicit scheme solves all of the equations simultaneously at
each grid point. The block-implicit formulation, generally requiring inversion of block-tridiagonal
matrices, entails more computational effort than a sequentially-implicit method. The additional
computation at each grid point and time step is usually compensated for by a dramatically improved
convergence rate. Block-implicit solvers can achieve CFL numbers in excess of 100, and often
converge in 100 to 200 time steps for flows including boundary-layer separation. For example, using
a block-implicit method, a supersonic two-dimensional shock-separated turbulent flow can be
simulated on an 80486-based microcomputer in about 3 hours of CPU time [Wilcox (1991)] . On the
same computer, a similar computation would take about 25 hours using a sequentially-implicit
method [Wilcox (1990)] and 75 hours using an explicit method [Wilcox (1974)]. As in the preceding
section, we begin with a brief overview of block implicit methods. For simplicity, we focus on a one-
dimensional system. The primary concern in this section is, of course, upon how turbulence model
source terms impact such methods. Now, suppose we choose to use a two-equation turbulence model
to determine the Reynolds stress. The following three points that must be considered in modifying a
block-implicit solution scheme.

106 Cambridge University Press


71

1. Decide whether to solve all equations simultaneously or to solve the model equations and
mean-flow equations sequentially.
2. If the preferred option is to solve all equations simultaneously, determine the changes to the
flux-Jacobian matrices.
3. Make provision for handling source terms.

In principle, solving all equations simultaneously will yield the most rapidly convergent scheme in
the number of iterations, but not necessarily in CPU time. However, the coupling between the
turbulence-model equations and the mean-flow equations appears to be relatively weak. The
primary coupling is through the diffusion terms, and the eddy viscosity is usually treated as a
constant in forming the viscous-flux Jacobian matrix. Limited experience to date seems to indicate
there is little advantage to solving all equations simultaneously as opposed to solving the model
equations and mean-flow equations sequentially. If all equations are solved simultaneously, the basic
system of equations for 1-D κ-ω model would be as follows:

Q 
 (F  Fv )  S
t x
 0 
 4 u   0 
ρ  ρu   μ  τ xx   
ρu   ρu 2  p  3 x  0 
  4 u  
    u  μ 
 τ xx   q̂ x  
 0 
Q  ρ E  , F  ( ρ E  p)u  , Fv    3 x   , S u 
 ρκ   ρ uκ   κ   τ xx x  β ρ ωκ 
     (μ  σ μ T )    ω  u 
ρ ω  ρ uω   x   
α   τ  β ρ ω 2

 (μ  σμ T ) ω    κ  xx x 
 x 

1 4 u 2
E  e  u 2  κ and τ xx  μ T  ρκ
2 3 x 3
I  F Fv  S  F Fv
  δx     ΔQ  δ x (F  Fv )  S , where
n n n
and are the Jacobian
 Δt  Q Q   Q  Q  Q
Eq. 5.1
Since dS/dQ is a diagonal matrix and its diagonal elements are always negative, its contribution is
guaranteed to enhance diagonal dominance of the matrix multiplying ΔQ.

View publication stats

You might also like