You are on page 1of 40

Applied Mathematical Modelling 39 (2015) 693–732

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

Review

Recent advances on the numerical modelling of turbulent flows


C.D. Argyropoulos a,⇑, N.C. Markatos b,c,⇑
a
Department of Chemical Engineering, Imperial College London, South Kensington Campus, London SW7 2AZ, UK
b
Computational Fluid Dynamics Unit, School of Chemical Engineering, National Technical University of Athens, 9 Iroon Polytechniou Str., Zografou Campus,
15780 Athens, Greece
c
Metropolitan College, School of Engineering, 74 Sorou Str., Marousi, Athens 15125, Greece

a r t i c l e i n f o a b s t r a c t

Article history: This paper reviews the problems and successes of computing turbulent flow. Most of the
Received 15 February 2013 flow phenomena that are important to modern technology involve turbulence. The review
Received in revised form 9 June 2014 is concerned with methods for turbulent flow computer predictions and their applications,
Accepted 7 July 2014
and describes several of them. These computational methods are aimed at simulating
Available online 14 July 2014
either as much detail of the turbulent motion as possible by current computer power or,
more commonly, its overall effect on the mean-flow behaviour. The methods are still being
Keywords:
developed and some of the most recent concepts involved are discussed.
Turbulence modelling
DNS
Some success has been achieved with two-equation models for relatively simple hydro-
LES dynamic phenomena; indeed, routine design work has been undertaken during the last
URANS three decades in several applications of engineering practise, for which extensive studies
DES have optimised these models.
Reynolds stress models Failures are still common for many applications particularly those that involve strong
curvature, intermittency, strong buoyancy influences, low-Reynolds-number effects, rapid
compression or expansion, strong swirl, and kinetically-influenced chemical reaction. New
conceptual developments are needed in these areas, probably along the lines of actually
calculating the principal manifestation of turbulence, e.g. intermittency. A start has been
made in this direction in the form of ‘multi-fluid’ models, and full simulations.
The turbulence modelling approaches presented here are, Reynolds-Averaged
Navier–Stokes (RANS), two-fluid models, Very Large Eddy Simulation (VLES), Unsteady
Reynolds-Averaged Navier–Stokes (URANS), Detached Eddy Simulation (DES) and some
interesting, relatively recent, hybrid LES/RANS techniques.
A large number of relatively recent studies are considered, together with reference to the
numerical experiments existing on the subject.
The authors hope that they provide the interested reader with most of the appropriate
sources of turbulence modelling, exhibiting either as much detail as it is possible, by means
of bibliography, or illustrating some of the most recent developments on the numerical
modelling of turbulent flows. Thus, the potential user has the appropriate information,
for him to select the suitable turbulence model for his own case of interest.
Ó 2014 Elsevier Inc. All rights reserved.

⇑ Current address: Metropolitan College, School of Engineering, 74 Sorou Str., Marousi, Athens 15125, Greece (N.C. Markatos). Tel./fax: +30 210 7723126.
E-mail addresses: c.argyropoulos09@imperial.ac.uk (C.D. Argyropoulos), n.markatos@ntua.gr (N.C. Markatos).
URL: http://www.amc.edu.gr (N.C. Markatos).

http://dx.doi.org/10.1016/j.apm.2014.07.001
0307-904X/Ó 2014 Elsevier Inc. All rights reserved.
694 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 694
2. Computer modelling of turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695
2.1. The differential equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695
2.2. Direct Numerical Simulation (DNS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 696
2.3. Reynolds-Averaged Navier Stokes (RANS) models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
2.3.1. Physical concepts of turbulence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
2.3.2. The equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
2.3.3. Zero-equation or algebraic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 699
2.3.4. Half-equation models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 700
2.3.5. One-equation models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 700
2.3.6. Two-equation models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 701
2.3.6.1 The k–e model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 701
2.3.6.2 Modified k–e model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 701
2.3.6.3 The k–x model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
2.3.6.4 More recent two-equation models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
2.3.6.5 Low Reynolds number modifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704
2.3.7. Non-Linear Eddy Viscosity Models (NLEVM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705
2.3.8. Recent advances in eddy viscosity modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 706
2.4. Differential Second-Moment (DSM) and Algebraic Stress Models (ASM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 707
2.5. Two-fluid models of turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 709
2.6. Large Eddy Simulation (LES) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 710
2.6.1. Validation of the LES approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 713
2.7. Monotone Integrated LES (MILES) and Implicit LES (ILES) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
2.8. Unsteady Reynolds-Averaged Navier–Stokes (URANS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
2.9. Very LES (VLES) and Detached-Eddy Simulation (DES). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 715
2.10. Hybrid RANS/LES strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 715
3. Applications of DNS and LES to flows in pipes and flows with a free surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 715
3.1. DNS of turbulent pipe flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 716
3.2. DNS of turbulent free-surface flows. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 719
3.3. LES of turbulent pipe flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 721
3.4. LES of turbulent free-surface flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 724
4. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 726
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 726

1. Introduction

Turbulence is the most complicated kind of fluid motion, making even its precise definition difficult. Literature contains
many definitions as, for example, that included in Markatos [1]: ‘‘A fluid motion is described as turbulent if it is three-dimen-
sional, rotational, intermittent, highly disordered, diffusive and dissipative’’.
Turbulence is a three-dimensional, time-dependent, nonlinear phenomenon. Its modelling is very attractive, as it saves
huge amounts of money, by avoiding the need to build and test prototypes, and as it transforms technologies by allowing
improved understanding of turbulence. This is particularly true in industrial flows which, apart from the complexities of
turbulence, involve also very complicated geometries and several design parameters, requiring optimisation [2]. Thus,
shape design is one of the most important drivers for the use of simulation approaches in fluid-engineering industry.
Examples refer to the drag of an aircraft or ship, propulsive efficiency of aeroengines or propellers, turbomachinery, chem-
ical process engineering, among others. In comparison to experiments, Computational Modelling offers a competitive
advantage if it is able to guide the analyst to a better design.
Computer programs now exist which are capable of solving three-dimensional, time-dependent Navier–Stokes (NS) equa-
tions, within practical computer resources. The reason that we do not make direct computer simulations of turbulence is that
turbulence is dissipated, and momentum exchanged by small-scale fluctuations [3]. The crucial difference between visuali-
sations of laminar and turbulent flows is the appearance of eddying motions of a wide range of length scales in turbulent
flows [4,5].
A typical flow domain of 0.1 m by 0.1 m with a high Reynolds number turbulent flow might contain eddies down to 10–
100 lm size. We would need computing meshes of 109 up to 1012 points to be able to describe processes at all length scales.
The fastest events take place with a frequency of the order of 10 kHz, so we would need to discretise time into steps of about
100 ls. We have estimated that the direct simulation of a turbulent channel flow at a Reynolds number of 800,000 requires a
computer which is half a million times faster than a current generation supercomputer. This estimate is analogous to the one
made by Speziale in 1991 [6], who stated that direct simulation of a turbulent pipe flow at a Reynolds number 500,000
required a computer 10 million times faster than the CRAY supercomputer of that time.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 695

With present day computing power it has only recently started to become possible to track the dynamics of eddies in
relatively simple flows at transitional Reynolds number. The computing requirements for the direct solution of the time-
dependent Navier–Stokes equations for fully turbulent practical flows at high Reynolds numbers are truly phenomenal
and must await major developments in computer hardware, possibly those based on quantum computing.
Meanwhile, engineers need computational procedures which can supply adequate information about the turbulent
processes, but which avoid the need to predict the effect of each and every eddy in the flow [7]. Therefore, in quantitative
work one is obliged to use turbulence models based on using averaged NS equations and, in addition, a set of equations
that supposedly express the relations between terms appearing in the NS equations [8–10]. It must be realised that most
of the available models pay no respect to the actual physical modes of turbulence (eddies, velocity patterns, high-vorticity
regions, large structures that stretch and engulf...) and, therefore, obscure the physical processes they purport to represent.
Flow visualisation experiments [11–16] confirm this point and demonstrate the difficulty of precise definition and
modelling. It is therefore hardly surprising that the actual physics of turbulence are nowhere to be seen in the available
models; simply because nobody can see as yet how mathematics can be employed to represent them in the models. It is,
however, also true that the engineering community has fortuitously often obtained very useful results by using relatively
simple models, such as those described in Section 2.3 below, results that would have required much more man-time and
experimental cost to obtain in their absence. Therefore, cautiously exercised and interpreted the turbulence models can be
valuable tools in research and design despite their physical deficiencies.
The purpose of the present effort is to provide a comprehensive review of the available turbulence modelling techniques.
The relevant material is certainly too much to be reviewed in a single paper. For this reason the authors confine attention to
what they consider the better established or more promising models. No disrespect is therefore implied for the models that
are scarcely – or not at all – mentioned. Extensive use has been made of the published literature on the topic and of earlier
reviews [17–20,11,21–40].
In addition, ERCOFTAC (European Research Community On Flow, Turbulence And Combustion) organises workshops and
special courses on best practise guidelines for CFD users. The Special Interest Group (SIG) 15 of ERCOFTAC is devoted to tur-
bulence modelling, and provides the appropriate data (e.g. experimental, DNS, highly-resolved LES databases) for the veri-
fication and validation of turbulence models, thus promoting their use for fundamental research and for industrial
applications [41].
Turbulent heat and mass transport are not explicitly covered in this review; the interested reader is directed to the review
by Launder [9]. Multi-phase phenomena are also not explicitly covered, apart from presenting the general differential equa-
tions and some necessary, to the authors’ mind, discussion on considerable work done for free-surface flows.
The review concludes with a summary of the advantages and disadvantages of the various turbulence models, in an
attempt to assist the potential user in choosing the most suitable model for his particular problem.
In the remainder of this review paper: Section 2 illustrates all of the available techniques for predicting turbulent flows;
in Sections 3, a literature survey is presented for applications using the LES and DNS technique; finally in Section 4, conclu-
sions and some recommendations for future research are outlined.

2. Computer modelling of turbulence

2.1. The differential equations

The motion of a fluid in three dimensions is described by a system of partial differential equations that represent math-
ematical statements of the conservation laws of physics (mass, momentum, energy and concentration conservation). The
momentum conservation equations are called the Navier–Stokes equations. In what follows the ‘‘Eulerian’’ equations gov-
erning the dynamics and heat/mass transfer of a turbulent fluid are given, in Cartesian tensor notation, using the
repeated-suffix summation convention. The equations are presented in the most general form of multi-phase flows
[42–54], as the single-phase ones are easily derived by just setting the volume fraction, ri equal to unity.
A convenient assumption for deriving these equations is based on the concepts of time- and space-averaging; it is that
more than one phase can exist at the same location at the same time [46,54]. Then, any small volume of the domain of inter-
est can be imagined as containing, at any particular time, a volume fraction ri of the ith phase. As a consequence, if there are n
phases in total,
X
n
r i ¼ 1: ð1Þ
i¼1

When flow properties are to be computed over finite time intervals, a suitable averaging over space and time must be carried
out. Following the above notion, that treats each phase as a continuum in the domain of interest, we can derive the following
balance equations:
Conservation of phase mass:
@
ðq ri Þ þ divðqi r i ~ _ i;
V iÞ ¼ m ð2Þ
@t i
696 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

where qi is the density, ~ _ i is the mass per unit volume entering


V i is the velocity vector, ri is the volume fraction of phase i; m
the phase i, from all sources per unit time, and div is the divergence operator (i.e. the limit of the outflow divided by the
volume as the volume tends to zero).
Summation of (2) over all phases leads to the ‘‘over-all’’ mass-conservation equation:
n 
X 
@
ðqi r i Þ þ divðqi r i ~
V i Þ ¼ 0; ð3Þ
i¼1
@t

which has of course a zero on the right-hand side.


Conservation of phase momentum:
@
V i uik Þ ¼ ri ð~
ðq ri uik Þ þ divðqi r i ~ k  grad p þ Bik Þ þ F ik þ lik ; ð4Þ
@t i
where: uik is the velocity component in the direction k of phase i; p is pressure, assumed to be shared between the phases; ~ k
is a unit vector in the k-direction; Bik is the k-direction body force per unit volume of phase i; Fik is the friction force exerted
on phase i by viscous action within that phase; and lik is the momentum transfer to phase i from interactions with other
phases occupying the same space.
Conservation of phase energy:
@
ðri ðqi hi  pÞÞ þ divðqi ri ~
V i hi Þ ¼ r i Q i þ H i þ J i ; ð5Þ
@t
where: hi is stagnation enthalpy of phase i per unit mass (i.e. the thermodynamic enthalpy plus the kinetic energy of the
phase plus any potential energy); Qi is the heat transfer to phase i per unit volume; Hi is heat transfer within the same phase,
e.g. by thermal conduction and viscous action; and Ji is the effect of interactions with other phases.
Conservation of species-in-phase mass:
@
ðq ri mil Þ þ divðqi r i ~ _ i M il ;
V i mil Þ ¼ divðri Cil grad mil Þ þ r i Ril þ m ð6Þ
@t i
where: mil is the mass fraction of chemical species l present in phase i; Ril is the rate of production of species l, by chemical
reaction, per unit volume of phase i present; Cil is the exchange coefficient of species l (diffusion); and Mil is the l-fraction of
the mass crossing the phase boundary, i.e. it represents the effect of interactions with other phases.
All of the above equations can be expressed in a single form as follows:
@
ðq ri ui Þ þ divðqi r i ~
V i ui Þ ¼ divðr i Cui grad ui Þ þ m
_ i Ui þ r i sui  total source of ui ; ð7Þ
@t i
where: ui is any extensive fluid property; the first term on the right-hand side expresses the whole of that part of the source
term which can be so expressed, with Cui being the exchange coefficient for ui. sui is the source/sink term for ui, per unit
phase volume; and m _ i Ui represents the contribution to the total source of any interactions between the phases, such as any
phase change (with Ui being the value of ui in the material crossing the phase boundary, during phase change). Distribution
of effects between m_ i Ui and sui is sometimes arbitrary, reflecting modelling convenience. For single-phase situations, the
above equations are valid by setting the r’s to unity.
For turbulent flow, averaging over times which are large compared with the fluctuation time leads to similar equations
for time-average values of ui with fluctuating-velocity effects usually represented by enlargement of Cui. More details on the
above concepts and equations may be found in [1].

2.2. Direct Numerical Simulation (DNS)

Solutions of turbulent flow problems (Eqs. (1)–(4)) can be obtained by using various analytical or numerical approaches,
with different level of accuracy in each case. Among the latter approaches, the Direct Numerical Simulation (DNS) has made a
significant contribution in turbulence research over the last decades [21], as it involves the numerical solution of the above
full three-dimensional, time-dependent Navier–Stokes equations without the need of any turbulence model. DNS is indeed
useful for the investigation of turbulence mechanisms, the improvement and development of turbulence models and for
assessing two-point closure theories.
Until the 1970’s the DNS approach was impossible to be used due to computer systems with insufficient memory and
speed to accommodate the required resolution needed for the small-scale turbulence effects. The first attempts for the inves-
tigation of homogeneous turbulence with DNS originated at the National Center for Atmospheric Research (NCAR) by Lilly
[55] and Orszag and Patterson [56] for 2-D and 3-D dimensional simulations, respectively. Rogallo [57] investigated the
effects of mean shear, rotational and irrotational strain on turbulence, based on the extension of Orszag and Patterson algo-
rithm. Kim et al. [58], Moser et al. [59], Abe et al. [60], Del Alamo et al. [61] and Hoyas and Jimenez [62], among others, per-
formed DNS for the investigation of wall turbulence for channel flows at1 Res = 180, 395, 640, 1900 and 2003.

1
Reynolds number based on the friction velocity us and the channel half width.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 697

There have also been extensive investigations of DNS in turbulent boundary layers at2 Reh = 700 [63], 1410 [64], 2500 [65],
2900 [66], 4060 [67] and 6650 [68,69]. More complex problems in wall-bounded turbulent flows (e.g. square duct, homoge-
neous isotropic turbulence, heat transfer, turbulence control and wavy boundary) have also been studied by DNS. The interested
reader may also consider the review papers by Moin and Mahesh [21], Kasagi and Shikazono [20] and Ishihara et al. [23]. The
most recent review concerning almost all the aspects of DNS (e.g. wall-bounded turbulence, turbulence control, bluff body tur-
bulence, turbulent flow structures and high performance computing) may be found in the work of Alfonsi [70].
The current largest scientific DNS was performed by Lee et al. [71] for turbulent channel flow at Res = 5200 and 3.5 times
more degrees of freedom than the DNS (40963 grid points) obtained by Kaneda et al. [72] and Kaneda and Ishihara [73] on
the Earth Simulator in Japan. The maximum Reynolds number obtained was approximately 1200 (Taylor microscale) which
is similar to the current capabilities obtained by laboratory experiments.
The absence of a turbulence model implies that the simulation is obtained by numerically solving over all the spatial and
temporal scales of turbulence, and its accuracy, therefore, is unrivalled by other methods. However, the DNS of high-Rey-
nolds number flows poses overwhelming demands on present-day available computing resources (speed and storage). It
is, therefore, necessary that DNS satisfies the following two constraints, according to Rogallo and Moin [17] and Kasagi
and Shikazono [20]:

(1) The dimensions of the computational domain must be large enough to comprise the largest turbulence scales.
(2) Grid resolution must be fine enough to capture the dissipation length scale, which is known as the Kolmogorov micro-
scale, g = (v3/e)1/4, where e is the average rate of dissipation of turbulence kinetic energy per unit mass, and v is the
kinematic viscosity of the fluid.

As a result, the required number of grid points for a given DNS is dependent on the Kolmogorov micro-scale and Kolmogo-
rov micro-timescale (s = (v/e)1/2) of the flow. The higher the Reynolds number, the finer the mesh should be. Hence, the
cell size in each direction of the computational domain should decrease with Re3/4 and the time step should decrease with
Re1/2 [74,3]. It is worth mentioning that the DNS time step is always smaller than the Kolmogorov micro-timescale in order
to maintain the algorithm’s numerical stability [75].
The required resolution for DNS in the directions parallel to the wall, according to the work of Kim et al. [58], is Dx+ = 8 and
Dz+ = 4, where Dx+ is the streamwise and Dz+ is the spanwise grid spacing, respectively. In wall-normal directions, a rule of
thumb is to place at least three grid points below y+ = 1 (non-dimensional distance from the wall to the first grid point) and at
least 10 grid points for y+ < 10, while in the outer region such as the pipe/channel centre line a value of Dy+ = 10 must be used.
Even with modern super-computers, the applicability of DNS is limited to flows of low to moderate Reynolds numbers.
Despite this current limitation, DNS is an effective and very useful tool for turbulence research leading to satisfactory results,
and used for testing simpler turbulence models, but it is still not practical for industrial or general engineering applications.
Among other benefits, DNS has contributed remarkably to testing conventional models and ideas and therefore to the devel-
opment of turbulence theory, in many ways which are summarized briefly below [20,74]. Furthermore, DNS data are impor-
tant for the development and improvement of turbulence models, due to the ability of DNS to provide the appropriate
turbulence statistics, including pressure and all spatial derivatives.
Important dimensionless numbers such as Reynolds and Prandtl can be varied in DNS, a fact of significant importance for
the derivation of a turbulence model with wide applicability. DNS is also suitable for studying a virtual flow which may occur
in reality now or in the near future. The latter advantage is important for the study of a dynamical turbulence phenomenon
[76] and for the evaluation of turbulence control methodologies [77].
Another important issue about DNS is the validation of the obtained results. According to Sandham [75] and Coleman and
Sandberg [78] the following are the criteria for such a validation: (a) validation of the obtained numerical data against
analytical solutions, experimental data and different numerical codes; (b) parametric studies with different grid resolutions,
domain sizes and time steps; (c) the time step (Dt) should be comparable with the Kolmogorov time-scale and the grid
spacing, Dxi, with Kolmogorov micro-scale, while the ratios of Dt/s and Dxi/s should be of order unity; (d) evaluation of
the statistical quantities budgets.

2.3. Reynolds-Averaged Navier Stokes (RANS) models

For the purpose of introducing the concepts of turbulent flow modelling we restrict attention to single-phase, incom-
pressible flow with constant laminar viscosity. The introduction of two-phase considerations, variable density and viscosity
are nowadays relatively easy tasks in modern solution algorithms. Only a generic presentation of turbulence modelling is
attempted here, for the sake of clarity. Details on the manner in which turbulence models properly couple into multi-phase
flow solvers may be found in literature (for example [79,80]).

2.3.1. Physical concepts of turbulence


Before discussing the turbulence models a very brief description of some concepts is provided. The main characteristic of
turbulence is the transfer of energy to smaller spatial scales across a continuous wave-number spectrum, i.e. a 3D, nonlinear

2
Reynolds number based on momentum thickness h and free-stream velocity.
698 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

process. A useful concept for discussing the main mechanisms of turbulence is that of an ‘eddy’ [81,82,3]. An eddy can be
thought of as a typical turbulence pattern, covering a range of wave- lengths, large and small eddies co-existing in the same
volume of fluid. The actual modes of turbulence are eddies and high-vorticity regions [4,3]. By analogy with molecular vis-
cosity, which is a property of the fluid, turbulence is often described by eddy viscosity as a local property of the fluid; the
corresponding mixing length in eddy viscosity models is treated in an analogous manner to the molecular mean-free path
derived from the kinetic theory of gases. This description is based on erroneous physical concepts but has proved useful in
the quantitative prediction of simple turbulent flows [7].
The eddies can be considered as a tangle of vortex elements (or lines) that are stretched in a preferred direction by mean
flow and in a random direction by one another. This mechanism, the so-called ‘vortex stretching’, ultimately leads to the
breaking down of large eddies into smaller ones. This process takes the form of an ‘energy cascade’. Since eddies of compa-
rable size can only exchange energy with one another [9], the kinetic energy from the mean motion is extracted from the
largest eddies [3]. This energy is then transferred to neighbouring eddies of smaller scales continuing to smaller and smaller
scales (larger and larger velocity gradients), the smallest scale being reached when the eddies lose energy by the direct action
of viscous stresses which finally convert it into internal thermal energy on the smallest-sized eddies [82]. It is important to
note that viscosity does not play any role in the stretching process nor does it determine the amount of dissipated energy; it
only determines the smallest scale at which dissipation takes place. It is the large eddies (comparable with the linear dimen-
sions of the flow domain), characterising the large-scale motion, that determine the rate at which the mean-flow kinetic
energy is fed into turbulent motion, and can be passed on to smaller scales and be finally dissipated. The larger eddies
are thus mainly responsible for the transport of momentum and heat, and hence need to be properly simulated in a turbu-
lence model. Because of direct interaction with the mean flow, the large-scale motion depends strongly on the boundary
conditions of the problem under consideration.
An increase in Reynolds number increases the width of the spectrum, i.e. the difference between the largest eddies (asso-
ciated with low-frequency fluctuations) and the smallest eddies (associated with high-frequency fluctuations). This suggests
that at high Reynolds numbers the turbulent motion can be well approximated by a three-level procedure, namely, a mean
motion, a large-scale motion and a small-scale motion [83].
Viscosity does not usually affect the larger-scale eddies which are primarily responsible for turbulent mixing, with the
exception of the ‘viscous sublayer’ very close to a solid surface. Furthermore, the effects of density fluctuations on turbulence
are small if, as in the majority of practical situations, the density fluctuations are small compared to the mean density, the
exception being the effect of temporal fluctuations and spatial gradients of density in a gravitational field. Therefore, one can
usually neglect the direct effect of viscosity and compressibility on turbulence. It is also important to note that it is the fluc-
tuating velocity field that drives the fluctuating scalar field, the effect of the latter on the former usually being negligible.

2.3.2. The equations


Eqs. (8)–(11) below constitute the mathematical representation of fluid flows, under the assumptions that the turbulent
fluid is a continuum, Newtonian in nature and that the flow can be described by the Navier–Stokes equations. For turbulent
flows, the latter represent the instantaneous values of the flow properties [1,84,85].
The equations for turbulence fluctuations are obtained by Reynolds de-composition which describes the turbulent motion
as a random variation about a mean value [1]:
 þ /0 ;
/¼/ ð8Þ
 its time- mean value and /0 the fluctuating part. The time-average of the fluc-
where / is the instantaneous scalar quantity, /
0
tuating value is zero / ¼ 0, and the mean value /  is defined as:
Z t 1 þDt
1
/ðxÞ ¼ limt!1 /ðx; tÞdt t 1  Dt  t 2 ; ð9Þ
Dt t1

where t1 is the time scale of the rapid fluctuations and t2 the time scale of the slow motion (for time-dependent mean value,
i.e. for non-stationary turbulence). By substituting Eq. (8) into the form of Eqs. (1)–(3) for single-phase, incompressible flows
and then taking the time-mean of the resulting equations, one derives the following continuity and NS equations:
i
@u
¼ 0; ð10Þ
@x

i
@u @ 
1 @p @2ui @
þ i u
ðu j Þ ¼  þm  ðu0 u0 Þ; ð11Þ
@t @xj q @xi @xj @xj @xj i j
where u i is the mean velocity, u0i the fluctuating velocity, q the fluid density and v the kinematic viscosity. Eq. (11) is known
as the Reynolds-Averaged Navier–Stokes (RANS) equation, while the term u0i u0j is the Reynolds-stress tensor:

sij ¼ u0i u0j ; ð12Þ

which is a symmetric tensor with six independent components. It is observed from Eqs. (10) and (11) that the number of
unknown quantities (pressure, three velocity components and six stresses) is larger than the number of the available
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 699

equations (continuity and Navier–Stokes). As a result, the system of equations is not yet closed and the problem of ‘‘closure’’
is reduced to the modelling of the Reynolds-stresses, in terms of mean-flow quantities.
The most popular approach to resolving this problem of ‘‘closure’’ is the use of the Boussinesq eddy-viscosity approxima-
tion [86]. The latter is based on an analogy between molecular and turbulent motions, in order to correlate Reynolds stresses
to the rate of strain of the mean motion. The turbulence eddies are thought of as parcels of fluid, which like molecules, collide
and exchange momentum, obeying the kinetic theory of gases. Thus, in analogy with the molecular viscous stress, the Rey-
nolds (turbulence) stresses are modelled as follows [1]:
 
2 @ui @ u j
sij ¼ u0i u0j ¼ jdij  mt þ ; ð13Þ
3 @xj @xi
1 1  02 
k ¼ u0i u0i ¼ u1 þ u02 02
2 þ u3 ; ð14Þ
2 2
where k is the turbulence kinetic energy and mt ð¼ lt =qÞ is the turbulence or eddy (kinematic) viscosity which, in contrast to
the molecular (kinematic) viscosity is not constant; and may vary significantly from flow to flow and from point to point [1];
and dij is the Kronecker delta. Substituting Eq. (13) into Eq. (11) leads to:
 
i
@u @ 
1 @p @ i
@u
þ i u
ðu j Þ ¼  þ ðm þ m t Þ : ð15Þ
@t @xj q @xi @xj @xj
The isotropic part of the Reynolds-stress tensor is absorbed normally into the pressure term as p ¼p  þ 2k
3
.
Dimensional analysis dictates that the unknown vt must be proportional to the product of a characteristic velocity Vt and
a characteristic length scale Lt. The difference between zero-equation, one-equation and two-equation models, discussed
below, lies in the way they choose to calculate them [1]. Thus, zero-equation models prescribe both characteristic velocity
and length-scale as algebraic expressions. One-equation models consider as characteristic velocity the square root of the tur-
bulence kinetic energy and prescribe algebraically the length scale, therefore:
pffiffiffi
vt ¼ C v 1 kL; ð16Þ
where C v 1 is a dimensionless constant. Two-equation models, such as k–e and k–x [1], described below in this subsection,
use differential equations to compute both the characteristic velocity and length scale and then estimate the value of vt by
the following equations:
( 2
C l f l ke ðk—e modelsÞ
vt ¼ ; ð17Þ
a xk ðk—x modelsÞ

where fl is a damping function, Cl and a are constants, e is the turbulence energy dissipation rate and x the dissipation per
unit turbulence kinetic energy.
Recent developments have led to the construction of non-linear eddy viscosity models, aiming at including non-linear
terms of the strain-rate [40]. More details for these models are presented in Section 2.3.7.
The traditional linear-eddy-viscosity RANS models may be divided into the following four main categories [1,87]: (a) alge-
braic (zero-equation) models, (b) half equation models (c) one-equation models and (d) two-equation models.
In the remainder of this section, a number of the better-established and most promising, according to the present authors’
experience, linear and non-linear eddy viscosity models, along with some more recently advanced ones, will be presented and
discussed.

2.3.3. Zero-equation or algebraic models


Zero-equation or algebraic models use partial differential equations only for computing the mean fields, while only alge-
braic expressions for the turbulence quantities [1]. This class of models is the oldest one, it is characterised by simplicity to
implement and has given good results for some applications of engineering relevance. For example, the best known of this
class, Prandtl’s mixing length model [88], is suitable for the prediction of thin-shear-layer flows such as boundary layers, jets,
mixing layers, and wakes. According to Prandtl, [88], in a boundary layer flow the eddy viscosity is given by:

@u
vt ¼ ‘2mix ; ð18Þ
@y
where ‘mix is the mixing length, that depends upon the type of flow, and is specified algebraically, while y is the direction
normal to the wall. This model is not suitable for predicting flows with recirculation and separation.
More modern variants of this category, following the contribution of Van Driest, Clauser and Klebanoff modifications [86],
are the Cebeci–Smith [89] and Baldwin–Lomax models [90]. These models are characterised by two-layer mixing-length
eddy viscosities, one as an inner and one as an outer layer viscosity. The second model is distinguished from the first because
of the different outer-layer length viscosity equation. Both are suitable for predicting turbulent flows in aerodynamics (e.g.
around airfoils) with similar accuracy, but are unreliable for separated flows. The mathematical formulations of these models
may be found in the textbook by Wilcox [86]. Nowadays, zero-equation models are used rarely and only for getting an initial
prediction of the flow field [91].
700 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

2.3.4. Half-equation models


In 1985 Johnson and King [92] developed a two layer model for the investigation of pressure driven separated flows. The
J–K model has been improved by Johnson [93] and Johnson and Coakley [94] in order to become applicable to compressible
flows as well. In addition, the J–K model has been extended for 3-D dimensional flows by Savill et al. [95]. The mathematical
equations along with comparisons against computational and experimental data may be found in Wilcox [86]. Even though
the J–K model has improved the classical algebraic models for predicting turbulent, transonic separate flows, it still suffers
from the same drawbacks as the Cebeci–Smith and Baldwin–Lomax models.

2.3.5. One-equation models


One-equation models are characterised by formulating one additional transport equation for the computation of a turbu-
lence quantity, usually the turbulence kinetic energy (k). For all of them there is still a need of prescribing a length-scale
distribution (L), which is defined algebraically and is usually based on available experimental data. For elliptic flows, like
recirculating and separated ones, experimental data is generally not available, making it difficult to prescribe algebraically
such a length scale. Therefore, most researchers decided to adopt two- or even more-equation models [1]. One-equation
models were used mainly in the nuclear and aeronautics industries (e.g. aircraft wings, fuselage, nuclear reactors) and the
most well known for aerospace applications are Baldwin and Barth [96] and Spalart and Allmaras [97] models. The Spal-
art–Allmaras was designed and optimised for flows past wings and airfoils and produced very good results. It is also easy
to implement for any type of grid (e.g. structured or unstructured, single-block or multi-block) [40]. However, both models
create enormous diffusion, in particular for regions of 3-D vortical flow [40]. Improvements of the aforementioned models
are presented in the works of Spalart and Shur [98], Dacles-Mariani et al. [99] and Rahman et al. [100], regarding the effects
of curvature, rotation, decrease of diffusion and for near-wall effects. Recent studies with the Spalart–Allmaras model have
been presented by Karabelas and Markatos [101] and Karabelas [102] for flow over an airfoil and past a flapping multi-ele-
ment airfoil, respectively. Karabelas [102] performed simulations past a plunging multi-element airfoil at Re = 6  105
(Fig. 1).
Accurate resolution of such flows still constitutes a great mathematical challenge for RANS modelling. It is well known
that the latter is often inaccurate even in terms of integrated quantities, such as lift and drag coefficients. This is due to

Fig. 1. Turbulence simulations of the flow past a plunging multi-element airfoil at Re = 6  105: Path-lines and pressure distribution at three fixed
geometric angles of attack, soaring flight regime (left), mid-time of the up-stroke (middle) and mid-time of the down-stroke phase (right), reproduced with
the author’s permission [102]. Reprinted by permission of the American Institute of Aeronautics and Astronautics, Inc.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 701

the fact that in flow regimes past multi-element airfoils, multiple transition processes from laminar to turbulent states and
vice versa could occur. In this study, we mention the one-equation model of Spalart Allmaras because its performance was
found to be superior [103] even over other more complete two-equation models. Further tuning of the latter models to
include transition effects did not increase considerably the accuracy of the simulations. It is worth mentioning, that in
the workshop held by NASA [104], for validating turbulence modelling of the flow past the multi-element airfoil MDA
30P-30N, mainly one-equation models were used in the governing equations.
Recently, Fares and Schroder [105] developed a complete and general one-equation model based on the two-equation k–
x model for predicting general turbulent flows such as wakes, jets, boundary layers and vortex flows. The new model proved
more accurate compared to the Spalart–Allmaras model, especially for jets and vortex flows. More information for zero-,
half- and one-equation models is provided in detail in the review papers by Markatos [1], Alfonsi [106] and in the classic
text by Wilcox [86].

2.3.6. Two-equation models


Two-equation models use, in addition to the mean-flow Navier–Stokes equations, two transport equations for two turbu-
lence properties. The first one is usually that for the turbulence kinetic energy (k) and the second any other from a variety
that includes: the dissipation rate of turbulence kinetic energy (e), the specific dissipation rate (x), the length scale (l), the
product of k  l, the time scale s, the product of k and s, among others [1]. This class of models is the most preferred by indus-
try it looks like remaining so for the foreseeable future [85]. Two-equation eddy viscosity models are still the first choice for
general CFD calculations, with the standard k–e model [107] and k–x [108] being the most widely used. There is no partic-
ular reason for this preference, but at least those models have been applied so widely, that we know their behaviour
beforehand.
In this section only the k–e and k–x models are presented, as being representative of the two-equation models, along
with their improvements and some interesting low-Re versions.

2.3.6.1. The k–e model


The k–e model is by far the most widely used and tested two-equation model, with many improvements incorporated
over the years. The standard k–e model of Launder and Sharma [107] is specified as follows:
Kinematic eddy viscosity (vt) equation:
2
k
mt ¼ C l : ð19Þ
e
Turbulence kinetic energy (k) equation:
 
@k @k @ ðm þ vt Þ @k i
@u
j
þu ¼  e þ sij : ð20Þ
@t @xj @xj rk @xj @xj
Turbulence dissipation rate (e) equation:
 
@e @e @ ðm þ vt Þ @ e e @ ui e2
j
þu ¼ þ C e1 sij  C e2 ; ð21Þ
@t @xj @xj re @xj k @xj k
where rk = 1.0 and re = 1.3 are the Prandtl numbers for k and e, respectively. The remaining model constants are: Cl = 0.09,
Ce1 = 1.44, Ce2 = 1.92. The standard k-e model behaves very in predicting turbulent shear flows, in many applications of engi-
neering interest. However, this model is unable to predict accurately flows with adverse pressure gradients and extra strains
(e.g. streamline curvature, skewing, rotation [91]). As a result it yields poor results for separated flows, whilst it is rather
difficult to be integrated through the viscous sublayer [86]. Despite the above shortcomings, the k–e model is recommended
for an at least gross estimation of the flow field and for cases such as combustion, multiphase flows and flows with chemical
reactions [91].

2.3.6.2. Modified k–e model


Improvements and modifications of the standard k–e model are many (for example, for flows with strong buoyancy, [1])
with probably the most important being the realisable k–e model [109] and the Renormalization Group (RNG) k–e model
[110].
The first model is based on the satisfaction of the realizability constraints on the normal Reynolds stresses and the
Schwartz inequality for turbulent shear stresses. Beside this, the Cl constant of standard k–e model is not anymore a
constant but it is computed in this improved model by an eddy-viscosity equation. Performance is substantially improved
for jets and mixing layers, channels, boundary layers and separated flows compared to the standard k-e model [109]. The
constants of the realisable k-e model are: Ce1 = 1.44, Ce2 = 1.9, rk = 1.0 and re = 1.2.
The RNG k–e model [110] is a modification of the classical k–e model with better predictions of the recirculation length in
separating flows. The model is represented by the same equations (19)–(21) of standard k–e model but with a modified coef-
ficient, Ce2, which is computed by the following equation:
C l g3 ð1  g=g0 Þ
C e2  C e2 þ ; ð22Þ
1 þ b1 g3
702 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

qffiffiffiffiffiffiffiffiffiffiffiffi  
Sk i @ u
1 @u j
g¼ ; S¼ 2Sij Sij ; S¼ þ ; ð23Þ
e 2 @xj @xi
where S denotes the mean strain-rate of the flow and Sij the deformation tensor. The model constants are: Cl = 0.085,
Ce1 = 1.42, Ce2 = 1.68, rk = re = 0.72, b = 0.012 and g0 = 4.38. This model gives better results than the standard k–e model
for separating flows, but fails to predict flows with acceleration [91].
Implementation of standard k–e model and RNG k–e model for pollutant dispersion from large tank-fires [111–114] and
street canyon flows [115], respectively, have been recently undertaken by the authors. In addition, high Reynolds number
turbulent flow past a rotating cylinder has also been examined by the authors and their colleagues. In Fig. 2, supercritical
streamline patterns are illustrated compared with laminar ones at Re = 200 and for some common examined cases (same
rotational rate) for flow past a rotating cylinder. More details for the study can be found in the work of Karabelas et al. [116].

2.3.6.3. The k–x model


Another ‘successful’ model and also widely used is the k–x model. The initial form of the model was proposed by
Kolmogorov in 1942 [117]. An improved version of the model was developed by the Imperial College group under Prof. B.
Spalding [118]. Further development and application of k–x model was performed by many scientists and engineers, but
the most important development was by Wilcox [108]. In the present paper, the most recent version of the model (Wilcox
(2006) k–x model) is presented below [86,108]:
Kinematic eddy viscosity (vt) equation:
( sffiffiffiffiffiffiffiffiffiffiffiffi)
k 2Sij Sij 7
mt ¼ ; x
~ ¼ max x; C lim ; C lim ¼ : ð24Þ
x
~ b 8

Turbulence kinetic energy (k) equation:


  
@k @k @ k @k i
@u
j
þu ¼ v þ r  b kx þ sij : ð25Þ
@t @xj @xj x @xj @xj
Specific dissipation rate (x) equation:
 
@x @x @ k@x rd @k @ x x @ ui
j
þu ¼ vþr  bx2 þ þ a sij : ð26Þ
@t @xj @xj x @xj x @xj @xj k @xj
The auxiliary relations and closure coefficients of the model are specified as follows:

a ¼ 0:52; b ¼ b0 f b ; b0 ¼ 0:0708; b ¼ 0:09; r ¼ 0:5; r ¼ 0:6; rd0 ¼ 0:125; ð27Þ


8 @k @ x
< 0;
> @xj @xj
60
1 þ 85vx Xij Xjk Ski

i @ u
1 @u j

rd ¼ @k @ x
> 0; fb ¼ ; vx  ; Xij ¼  ; ð28Þ
> rd0;
: @xj @xj 1 þ 100vx ðb xÞ
3 2 @xj @xi

where Clim is the stress-limiter strength, fb the vortex-stretching function, vx the dimensionless vortex-stretching parameter
and Xij the mean-rotation tensor. The k–x model is superior to the standard k–e model for several reasons. For instance, it
achieves higher accuracy for boundary layers with adverse pressure gradient and can be easily integrated into the viscous
sub-layer without any additional damping functions [86]. In addition, the recent version of Wilcox (2006) k–x model is
much more accurate for free shear flows and separated flows. The model still suffers from weaknesses when applied to flows
with free-stream boundaries (e.g. jets), according to the review paper by Menter [119].

2.3.6.4. More recent two-equation models


A more advanced turbulence model is the Shear Stress Transport (SST) model by Menter [120]. This model combines the
advantages of k–e and k–x models in predicting aerodynamic flows, and in particular in predicting boundary layers under
strong adverse pressure gradients. The model has been validated against many other applications with good results such as
turbomachinery blades, wind turbines, free shear layers, zero pressure gradient and adverse pressure gradient boundary
layers. Recent improvements of the model are an enhanced version for rotation and streamline curvature [121] and the
replacement of the vorticity in the eddy viscosity with the strain rate [119]. The mathematical formulation of the model
is not repeated due to space limitations, but it may be found in the above mentioned references.
Another class of two-equation models is the two-time scale models, with significantly improved results compared to the
k–e model. Hanjalic et al. [122] proposed a multi-scale model in which separate transport equations are solved for the
turbulence energy transfer rate across the spectrum. The mathematical formulation of the proposed turbulence model is
as follows [123].
 
kkP
vt ¼ C l ; ð29Þ
eP
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 703

Fig. 2. Streamline patterns at Re = 200, 5  105, 106, 5  106 and rotational rates a = 2, 3, 4, 5, 6, 7 and 8. L1 and L2 stagnation points are apparent for low
rotational rates (laminar flow), while A, B, C and Z are addressed for the super-critical Reynolds numbers [116].
704 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

 
DkP @ @kP
¼ ðm þ vt Þ þ Pk  eP ; ð30Þ
Dt @xj @xj
 
DkT @ @kT
¼ ðv þ vt Þ þ eP  eT ; ð31Þ
Dt @xj @xj
 
DeP @ @ eP eP e2 l @ u
@u i
¼ ðv þ vt Þ þ C P1 Pk  C P2 P þ C 0P1 kP elmk eijk ; ð32Þ
Dt @xj @xj kP kP @xm @xj
 
DeT @ @ eT eP eT eT
¼ ðv þ vt Þ þ C T1  C T2 2 : ð33Þ
Dt @xj @xj kT kT
The form of Eq. (29) has been obtained from simplifying the mean-Reynolds-stress (MRS) equation by considering normal
Reynolds stresses proportional to k and by taking the time scale for pressure strain to be kePP . The above mentioned Eqs.
(29)–(33) use the following set of coefficients and functions:
kP
k
1 eP
C l ¼ 0:09; C P1 ¼ 2:2; C 0P1 ¼ 0:11; C P2 ¼ 1:8  0:3 kT ; C T1 ¼ 1:08 ; C ¼ 1:15; ð34Þ
P
kT
þ1 eT T2

where P k ¼ ui uj @@xui . Here kp and kI are, respectively, the turbulence kinetic energy in the production and dissipation ranges,
j

Pk is the rate at which turbulence energy is produced (or extracted) from the mean motion, ep is the rate at which energy is
transferred out of the production range, eT is the rate at which energy is transferred into the dissipation range from the iner-
tia range and e is the rate at which turbulence energy is dissipated (i.e. converted into internal energy).
It is worth mentioning that the proposed version of k–e model performs better than the standard (single-scale) k–e model
due to the fact that ep (rate at which energy is transferred out of the production range) replaces Pk (turbulence production) in
the dissipation rate (e) equation, simply because in flows where Pk is suddenly switched off, e is not expected to decrease
immediately. The present model gives better predictions than the single-scale k–e model in plane and round jets [1].
The main advantage of the two-scale k–e model is the combination of modelling the cascade process of turbulence kinetic
energy and of solving complex flows such as separating and reattaching flows. Improvements of the model may be achieved
by accounting for the proper empirical coefficients which affect the spectrum shape. Applications of the model for breaking
waves [124], plane synthetic and swirling jet [125], wake-boundary-layer interaction and compressible flow [126], may be
found in the literature. Finally, new models for non-equilibrium flows have also been developed by Klein et al. [127] with
satisfactory results.

2.3.6.5. Low Reynolds number modifications


Most of the above models are applicable for turbulent flows at high-Re numbers, but are inaccurate for the prediction of
the flow in the vicinity of the wall, where viscous forces dominate. In order to treat this shortcoming, many scientists and
engineers have proposed a number of near-wall modifications. These models with near-wall modifications are referred to as
‘‘Low-Reynolds Number’’ (LRN) models. A full list of these models is presented in the text by Wilcox [86] and in the review
paper by Patel et al. [128]. In the present paper two popular LRN models will be presented, the Lam–Bremhorst k–e model
[129] and Bredberg et al. k–x model [130]. The mathematical formulation of the first model can be written in the following
boundary layer form:
    2
@k @k @ vt @k @u

u þt
 ¼ vþ þ vt  e; ð35Þ
@x @y @y rk @y @y
    2
@ ~e @ ~e @ vt @ ~e ~e 
@u ~e2

u þt
 ¼ vþ þ C e1 f 1 v t  C e2 f 2 þ E; ð36Þ
@x @y @y rk @y k @y k
where the turbulence dissipation (e) is given by the following equation:
e ¼ e0 þ ~e ð37Þ
and e0 is the value of e at y = 0. The kinematic eddy viscosity is defined as:
2
k
vt ¼ C l f l : ð38Þ
~e
The damping functions ðf 1 ; f 2 ; f l ; e0 and EÞ and closure coefficients for the Lam–Bremhorst k–e model are presented below:
  !3 2 0:5
20:5
2 0:05 k k y
f l ¼ ð1  expð0:0165Ry ÞÞ 1 þ ; f1 ¼ 1 þ ; f 2 ¼ 1  expðRe2t Þ; Ret ¼ ; Ry ¼ ;
Ret fl ~ev v
e0 ¼ 0; E ¼ 0; C e1 ¼ 1:44; C e2 ¼ 1:92; C l ¼ 0:09; rk ¼ 1:0; re ¼ 1:3: ð39Þ
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 705

The development of LRN k–e models improved the original k–e model by making it more compatible with the law of the
wall. However, LRN modifications did not improve the problem with strong adverse pressure gradient. More details for the
difficulty of LRN k–e models to predict turbulent flows with pressure gradients may be found in Wilcox [86]. Finally, Patel
et al. [128] claim that any improvement in predicting flows with adverse pressure gradient would require modifications to
the original k–e model itself.
Another class of LRN models is the k–x models and one of the most popular is the standard k–x model by Wilcox [108].
Extensions and improvements of the model have been proposed by Wilcox [131], Peng et al. [132], Bredberg et al. [130],
among others.
The mathematical equations of Bredberg et al. k–x model are as follows:
  
@k @ @ vt @k
þ  j kÞ ¼
ðu vþ þ Pk  C k kx; ð40Þ
@t @xj @xj rk @xj
   v v  @k @ x
@x @ @ vt @x t x
þ  j xÞ ¼
ðu vþ þ Cx þ þ C x 1 P k  C x 2 x2 : ð41Þ
@t @xj @xj rx @xj k k @xj @xj k
The turbulence kinematic viscosity is defined as:
k
vt ¼ C l f l ; ð42Þ
x
where the damp function fl is given by the following equation:
!" (   )#
2:75
1 Ret
f l ¼ 0:09 þ 0:91 þ 1  exp  : ð43Þ
Re3t 25

Finally, the constants of the model are denoted as:


C l ¼ 1; C k ¼ 0:09; C x ¼ 1:1; C x1 ¼ 0:49; C x2 ¼ 0:071; rk ¼ 1; rx ¼ 1:8: ð44Þ
The model of Bredberg et al. [130] presents improved results against the original Wilcox k–x model, compared to DNS
and experimental data for three different cases (channel flow, backward facing step flow and rib-roughened channel flow).
Recently, an extension of the model to viscoelastic fluids was proposed by Resende et al. [133].

2.3.7. Non-Linear Eddy Viscosity Models (NLEVM)


As mentioned earlier in Section 2.3.2, the Non-Linear Eddy Viscosity Models (NLEVM), may be defined as non-linear
extensions of the eddy-viscosity models in which Eqs. (13) and (15) can be rewritten in a more general form, in order to
include non-linear terms of the strain-rate [40]:

2 X
N
ðnÞ
sij ¼ u0i u0j ¼ jdij þ an T ij ; ð45Þ
3 nþ1

 
i
@u @ 
1 @p @ i
@u
þ i u
ðu j Þ ¼  þ ðm þ m t Þ þ N:S:T:; ð46Þ
@t @xj q @xi @xj @xj
where N.S.T are non-linear source terms deriving from Eq. (45).
This class of models has been developed to overcoming the deficiencies of eddy-viscosity models, in particular for two-
equation models. There is a large number of NLEVM in the literature and they may be categorised as quadratic and cubic
models. Popular quadratic models have been proposed by Gatski and Speziale [134] and Shin et al. [135], among others.
The first model is a high-Re k–e model which supports separation in adverse pressure gradient flows. The model of Shin
et al. [135] has shown improved results for backward facing step compared to classical linear eddy viscosity models, but
it also suffers with rotational effects especially for channel flow [136].
The mathematical formulation of the cubic model by Craft et al. [137] is selected for presentation here, as a general form
of the category. The anisotropic tensor and turbulence kinetic energy are defined as:
ui uj 2 1
aij ¼  dij and k ¼ uk uk : ð47Þ
k 3 2
The mathematical formulation of the cubic model is as follows [137]:
   
vt vt 1 vt vt 1 vt k
aij ¼  Sij þ c1 Sik Skj  Skl Skl dij þ c2 ðXik Skj þ Xjk Skl Þ þ c3 Xik Xjk  Xlk Xlk dij þ c4 2 ðSki Xlj
k ~e 3 ~e ~e 3 ~e
 
vt k 2 vt k vt k
þ Skj Xli ÞSkl þ c5 2 Xil Xlm Smj þ Sil Xlm Xmj  Slm Xmn Xnl dij þ c6 2 Sij Skl Skl þ c7 2 Sij Xkl Xkl ; ð48Þ
~e 3 ~e ~e

where Sij is the mean strain-rate tensor and Xij the mean vorticity tensor:
706 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Table 1
The model coefficients [136].

c1 c2 c3 c4 c5 c6 c7
f
0:05 f q
f
0:11 f q f ~S 0.8fc 0 0.5fc 0.5fc
l l
0:21 f ð~Sþq X~ Þ=2
l
Cl fl
p~ffie
0:667r n f1exp ½0:415 expð1:3n5=6 Þ g ~
1:1 e½10:8 expðRt =30Þ
1þ1:8n 1þ0:6A2 þ0:2A3:5
2

rn fq fc
n o rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ~  rn r 2n
1 þ 1  exp½ð2A2 Þ3 Rt 0:5
1 þ 4 exp  20 ð1þ0:0086g2 Þ 1þ0:45n2:5

i @ u
@u j i @ u
@u j
Sij ¼ þ Xij ¼  : ð49Þ
@xj @xi @xj @xi
The empirical coefficients of the model are presented in Table 1, where ~ ~ the dimen-
S is the dimensionless strain parameter, X
sionless vorticity parameter, ~e the homogenous dissipation rate, which are defined as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi k qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
~S ¼ k Sij Sij =2; ~¼
X Xij Xij =2; vt ¼ c l
k
: ð50Þ
~e ~e ~e
The model appears better compared to ordinary linear eddy viscosity models (e.g. for impinging jet flows). The compu-
tational time is approximately 20% more compared to a low-Re k–e model. One drawback of the model is its performance for
convex surfaces, according to Craft et al. [137].
Another popular cubic low-Re k–e model was developed by Apsley and Leschziner [138], with its free parameters cali-
brated with data from DNS data for channel flow. The model leads to better results for airfoil and diffuser flows compared
to other linear and nonlinear EVM. For more details for the NLEVM, the interest reader is directed to the review paper by
Hellsten and Wallin [136].

2.3.8. Recent advances in eddy viscosity modelling


It is worth mentioning some recent eddy-viscosity models, such as Durbin’s t2-f model (also known as v2–f) [139] and f–f
model [140]. The t2–f model is based on the elliptic relaxation concept and employs two additional equations, apart from the
k and e ones. One for the velocity scale t2 and one for the elliptic relaxation function, f. The main motivation for the devel-
opment of this model was the improved modelling in the vicinity of the wall (near-wall turbulence). More applications (e.g.
rotating cylinder, rotating channel flow, axially rotating pipe and square duct) and validation of the model with experimental
and DNS data, may be found in the work of Durbin and Petterson [141].
The f–f model is based on the similar concept of elliptic relaxation, but instead of solving the t2 equation it solves for the
velocity scale ratio f = t2/k [140]. The full equations of the f–f model, which are similar to those of the t2–f model, are [140]:
vt ¼ C l fks; ð51Þ
  
Dk @ vt @k
¼ vþ þ P  e; ð52Þ
Dt @xj rt @xj
  
De @ vt @ e ðC e1 P  C e2 eÞ
¼ vþ þ ; ð53Þ
Dt @xj re @xj s
  
1 P 2
L2 r2 f  f ¼ c1 þ C 02 f ; ð54Þ
s e 3
  
Df @ vt @f f
¼ vþ þ f  P: ð55Þ
Dt @xk rf @xk k
Completeness of the model is achieved by Durbin’s [142] realizability constraints, combined with the lower bounds (Kol-
mogorov time- and length- scale):
" ! #
a v0:5
k
s ¼ max min ; p ffiffiffi ;C ; ð56Þ
e 6C l jSjf s e
" !  3 0:25 #
1:5 0:5
k k v
L ¼ C L max min ; pffiffiffi ; Cg ; ð57Þ
e 6C l jSjf e
where a 6 1 (recommended a = 0.6 [140]). The coefficients of this model are: Cl = 0.22, Ce1 = 1.4(1 + 0.012/f), Ce2 = 1.9,
c1 = 0.4, C 02 ¼ 0:65, rk = 1, re = 1.3, rf = 1.2, Cs = 6.0, CL = 0.36, Cg = 85.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 707

Fig. 3. Mean velocity streamlines over the swept wing surface, highlighting regions of interest [145].

It should be noted that the f–f model is more stable [141] compared to the t2–f model. Both models are better for com-
puting wall-bounded flows compared to the classical low-Re two equation models (e.g. k–x and k–e), but they are still weak
against DSM (introduced next in Section 2.4) and advanced NLEVMs. The f–f model has been extensively validated with
experimental and DNS data for plane channel, backward-facing step and multiple-impinging jets flows, presenting satisfac-
tory agreement.
Recently, a new robust version of the t2–f model was proposed by Billiard and Laurence [143] with improved numerical
stability, known as the BL-t2 =k. The model is based on the elliptic blending method of Manceau and Hanjalic [144] and was
validated for pressure induced separated flows, as well as buoyancy impairing turbulent flows, with satisfactory results.
More detailed evaluation of the model against other turbulence models and test cases (e.g. 3-D diffuser and swept wing)
is presented in the work of Billiard et al. [145]. In Fig. 3, the mean velocity streamlines for swept wing are compared to data
from Implicit Large Eddy Simulation (LES). Both models capture the leading edge vortex but the secondary vortex region
(red-dashed line) is reproduced by the EBRSM model and secondly from the BL-t2 =k model. A detailed review of t2–f model
evolution may be found in the work of Billiard and Laurence [143].

2.4. Differential Second-Moment (DSM) and Algebraic Stress Models (ASM)

A type of turbulence closure models with great expectations to replace the widely used k–e model is the Differential
Second-Moment (DSM) or Reynolds Stress (RS) or Mean-Reynolds Stress (MRS) model. DSM presents natural superiority
compared to the two equations turbulence models, as it is physically the more complete model (history, transport and
anisotropy of turbulent stresses are all accounted for). More specifically, DSM closure models explicitly employ transport
equations for the individual Reynolds Stresses, u0i u0j (as well as for u0j T 0 ), each of them representing a separate velocity scale.
The transport equation of u0i u0j for an incompressible fluid, excluding effects of rotation and body force, may be written in
general symbolic tensor form as follows:
Lij þ C ij ¼ Pij þ /ij þ Dij  eij ; ð58Þ

where Lij is the local change in time, Cij the convective transport, Pij the production by mean-flow deformation, /ij the stress
redistribution tensor due to pressure strain, Dij the diffusive transport and eij the viscous dissipation tensor. The Lij, Cij and Pij
terms do not require any modelling and are given by the following equations:
 
@ @
Lij þ C ij ¼ m
þu u0 u0 ; ð59Þ
@t @xm i j
 
j
@u i
@u
Pij ¼  u0i u0m þ u0j u0m : ð60Þ
@xm @xm
The remaining terms, /ij , Dij and eij need to be modelled. The simplest way to model the viscous dissipation tensor is by
assuming local isotropy:

2 @um @um
eij ¼ edij ; e ¼ v ; ð61Þ
3 @uk @uk
where e is the turbulence dissipation and dij the Kronecker unit tensor. The diffusion of turbulence (Dij) is usually treated by
the popular Daly–Harlow model [146]:
708 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

!
0 0
@ k @ui uj
Dij ¼ C s u0l u0m with C s ¼ 0:25: ð62Þ
@xk e @xl

Instead of the popular Daly–Harlow model, there are more advanced models that can be used and have been developed
over the years, such as the models by Magnaudet [147] and Nagano and Tagawa [148]. More details and validation of the
above mentioned models can be found in the review paper by Hanjalic [149].
The radically new feature of the DSM-equation is the pressure-strain ‘redistribution’ term (/ij ), which does not appear in
the exact solution of k–e equation. This suggests that the pressure-strain term only serves to redistribute the turbulence
energy among its components and to reduce the shear stresses, thus tending to make the turbulence more isotropic. The
unknown correlations appearing in the DSM-equation are either determined by a transport equation or else they are
expressed in terms of second-order correlations ðu0i u0j Þ themselves; the latter procedure, closing the DSM-equation at its
own level, is often referred to as ‘second-order closure’. The redistribution term, /ij , is usually modelled by the Isotropization
of Production (IP) model [150]:

   
2 e dij
/ij ¼ C 1 u0i u0j  dij k  C 2 Pij  P kk with C 1 ¼ 1:8 and C 2 ¼ 0:6: ð63Þ
3 k 3
Improvements of the IP pressure-strain model are the LRR-QI model by Launder et al. [151], the SSG model by Speziale
et al. [152], the CL model by Craft and Launder [153], and LT model by Launder and Tselepidakis [154]. For detailed evalu-
ation of the models, the interest reader is referred to the work of Hanjalic and Jakirlic [155]. DSM closure models predict
more accurate physical phenomena which involve streamline curvature, strong pressure gradients, swirling and system rota-
tion effects [155].
It is important to mention that the initial versions of DSM models could not perform very well in handling the return to
isotropy [1]. They may, however, work well in flows dominated by other effects. In addition, the DSM models do not always
perform better than the two-equation models. For instance, recent numerical simulations in street canyon flows performed
by Koutsourakis et al. [115] indicated that DSM performance was not good enough according to the theoretical expectations,
while the RNG k–e model exhibited better results. The evaluation of the models (DSM, RNG k–e and standard k–e) was done
comparing to different experimental and numerical data (LES). Figs. 4 and 5 present numerical results for the velocity pro-
files compared with experimental data. It is concluded that DSM, at least as it has been applied in that study, is not more
useful than simpler models for practical use. However, both modelling and experimental uncertainties are high, so that extra
attention is required in order to draw any definitive conclusions about the quality of the models.
Main disadvantages of DSM are the difficulty in the modelling of more terms in the turbulence equations and the
increased demand on computer resources. The new generation of DSM closure models have solved most of the above-men-
tioned difficulties but the computational demands are roughly twice as large as those for the two-equation models, for high -
Re number flows using wall functions [155].
It is worth noting the Elliptic Blending DSM by Manceau and Hanjalic [144], which belongs to the category of advanced
DSMs. The model is based on the DSM of Durbin [159] but, instead of resolving equations for the stress components, it adopts
a single elliptic equation [160]. Implementation of EBDSM for predicting impinging jet flows can be found in the work of
Thielen et al. [161].
Another interesting approach is the hybridization of DSM with an eddy viscosity model, recently developed by Basara and
Jakirlic [162]. The model combines the advantages of DSM along with the robustness of k–e model and it is known as Hybrid
Turbulence Model (HTM). It may be used as an initialization model for DSMs in order to stabilize it and reduce the

Fig. 4. Experiment 4: Non-dimensional horizontal velocity profiles at leeward and windward side of a street canyon with aspect ratio H/D = 1 and very
rough walls. Uref is the free stream velocity, equal to 8 m/s. Experimental data are extracted from Kovar-Panskus et al. [156]. CFD results with the k-e model
from the original paper are also included [115]. ‘‘N. Koutsourakis, J. Bartzis, N. Markatos, Evaluation of Reynolds stress, k-e and RNG k-e turbulence models in
street canyon flows using various experimental datasets, Environmental Fluid Mechanics 12 (2012) 379–403. This is Fig. 6 in the publication in which the
material was originally published. With kind permission from Springer Science and Business Media.’’
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 709

Fig. 5. Experiment 6: Performance of the three turbulence models against experimental data extracted from Li et al. [157] in a street canyon with aspect
ratio H/D = 2. LES results of Li et al. [158] are also included [115]. ‘‘N. Koutsourakis, J. Bartzis, N. Markatos, Evaluation of Reynolds stress, k-e and RNG k-e
turbulence models in street canyon flows using various experimental datasets, Environmental Fluid Mechanics 12 (2012) 379–403. This is Fig. 8 in the
publication in which the material was originally published. With kind permission from Springer Science and Business Media.’’

computational time. The model has been tested for many flows such as flow around a car, axially rotating pipe, 180° turned U
bend, backward-facing step and round jet impinging on a flat plane. The results were better than EVMs and closer to those
obtained by full DSM.
Simplified versions of DSMs are the Algebraic Stress Models (ASM) or Explicit Algebraic Reynolds Stress Models (EARSM)
[136], obtained by eliminating the transport terms, using instead those of the kinetic energy equation. EARSM have the rep-
utation of being simple and easy to implement for boundary layer flows, while for elliptic, recirculating flows they are very
unstable. In addition, their performance is dependent on the DSM from which they were derived. This category constitutes
an intermediate-level between DSM and eddy viscosity models. EARSM are characterised by less computational demands
and higher accuracy compared to LEVM.
First attempt for the derivation of EARSM was done by Pope [163] for two-dimensional flows and it was later extended
and refined by Gatski and Speziale [134] and by Jongen and Gatski [164] for 3-D flows. The EARSM are popular for predicting
aeronautical flows, in particular with the model by Wallin and Johansson [165].
Due to space limitations, the interested reader is directed to the recent review papers by Hellsten and Wallin [136] and
Alfonsi [106], for the mathematical equations of EARSM.

2.5. Two-fluid models of turbulence

Undoubtedly, the main idea behind a large number of turbulence models derives from the notion of Boussinesq, who
introduced the idea of an effective viscosity, and of Prandtl who conceived the notion of turbulence mixing phenomena being
very similar to those treated by the dynamical theory of gases. The mathematical background of convectional turbulence
theory reflects only the unstructured diffusion of the molecular-collision process; large structure formation and growth,
and fine-structure creation and stretching, are nowhere to be found.
The above-mentioned facts led Spalding [166], among others, to the development of the ‘two-fluid’ theory, briefly pre-
sented below. The origins of two-fluid model ideas are to be found back in the 1940’s and 1950’s, particularly in the field
of turbulent combustion.
The following two-fluid model results in the prediction of the intermittency and of the conditional flow variables within
the turbulent and non-turbulent zones of the flow. The model was proposed by Spalding [166] and then developed by Malin
[167] for the investigation of intermittency in free turbulent shear flows.
The Spalding model is designed on an analogy between intermittent flows and two-phase flows, rather than on any spe-
cific and rigorous closure of conditional-averaged transport equations. It is supposed that two-fluids share occupancy of the
same space, although not necessarily at the same time, their share of space being measured by the volume fractions. There
are many ways in which the two-fluids can be distinguished. In case of turbulence intermittency computations, it is conve-
nient to define the fluids as ‘turbulent’ and ‘non-turbulent’. The equations governing the motion of the turbulent and non-
turbulent fluids are given in detail by Markatos [1,79].
Improvement and expansion of the model for turbulent combustion was presented by Markatos and Kotsifaki [168]. Fur-
thermore Shen et al. [169] used a two-fluid model to simulate turbulent stratified flows, while Yu et al. [170] established a
modified two-fluid model to simulate the flow and heat transfer characteristics of air curtains in an open vertical display
cabinet. For Rayleigh–Taylor mixing so far only two models have been developed, namely Youngs’ model [171] and two-
structure two-fluid two-turbulent (2SFK) model [172]. Liu et al. [173] applied two-fluid model for predicting the flow in
UV disinfection reactor. Finally, Cao et al. [174] used the model for the design of air curtains for open vertical refrigerated
display cases. The numerical results compared with experimental data present good agreement.
710 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Fig. 6. Turbulence energy spectrum. The figure has been redrawn based on [175].

2.6. Large Eddy Simulation (LES)

Another modelling approach, promising to be more accurate and of wider applicability than RANS and less computation-
ally demanding than DNS, is the Large Eddy Simulation (LES) approach. In LES of turbulence, the important large scales are
fully resolved whilst the small sub-grid scales are modelled. The main advantage of LES compared to RANS models is that in
the former only the small, isotropic turbulent scales are modelled and not the entire spectrum (Fig. 6) as it is the case in the
latter. The LES approach is extremely useful for the investigation of turbulence at high Reynolds numbers, for the develop-
ment and assessment of new turbulence models, and for the prediction of complex flows where other turbulence models
may prove inadequate [176,177].
The first attempts of LES are found in the pioneering works of Smagorinsky [178] and Lilly [179] in meteorology and Dear-
dorff [180] in engineering. Since then, LES has seen a tremendous popularity for the study of turbulent flows. The develop-
ment and testing of LES has concentrated at first on isotropic turbulence by Kraichnan [181] and Chasnov [182] and on
turbulent channel flow by Deardorff [180], Schumann [183], Moin and Kim [184] and Piomelli [185]. The basic steps of
LES method, according to Pope [24] and Berselli et al. [25] are: (a) a filtering operation of the N–S equations; (b) a closure
model for SGS stress tensor; (c) imposition of the appropriate boundary and initial conditions with special care for near wall
modelling; (d) selection of the suitable numerical method for spatial and temporal discretization of N–S equations and (e)
performance of the numerical simulation.
Instead of time-averaging, a spatial filtering approach is adopted in order to separate the resolved (large-eddy) field from
the small-eddy (sub-grid) field. A filter operation is defined by the convolution [186]:
Z

/ðxÞ ¼  Þdx0 ;
/ðx0 ÞGðx; x0 ; D ð64Þ

 is the filter width. The filter function is respon-


where D is the domain of integration, G is the specified filter function, and D
sible for determining the structure and size of the small scales that are captured [32]. More details about spatial filtering
techniques are presented in Aldama [187], Pope [24] and Sagaut [26]. An illustrative example of LES for single-phase, incom-
pressible flow is given below.
Assuming that the filtering operation presented above commutes with temporal and spatial derivatives, the filtered gov-
erning equations can be expressed as follows:
Filtered continuity equation:
i
@u
¼ 0: ð65Þ
@xi
Filtered Navier–Stokes equations:
i @ u
@u i u
j  @ sij
1 @p @2ui
þ ¼  þv ; ð66Þ
@t @xj q @xi @xj @xj @xj
where sij denotes the Sub-Grid Scale (SGS) stress tensor, defined as:
sij ¼ ui uj  ui uj : ð67Þ
The SGS stress tensor is symmetric, invariant to Galilean transformation and plays important role for the dynamic cou-
pling between the small and large scales of turbulence.
Denoting the fluctuation /0 of / with respect to / as:
 ¼ /  /0 ;
/ ð68Þ
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 711

the SGS stress tensor can be written as:

sij ¼ ui uj  ui uj ¼ ui uj  ui uj þ ui u0j þ u0i uj þ u0i u0j ; ð69Þ
|fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} |{z}
Lij C ij Rij

where Lij is the ‘‘Leonard term’’, C ij is the ‘‘Cross term’’, and Rij the ‘‘Reynolds term’’ according to the work of Clark et al. [188].
Each of these terms describes physical interactions between the scales that arise from the LES approach. More specifically,
the Leonard term, Lij , expresses the interactions between resolved scales and it is an explicit term that can be computed from
the filtered field without a modelling approach. Subsequently the use of a spectral or sharp cut-off-filter transforms the
Leonard terms in aliasing error [32]. Moreover, the Cross term, C ij , represents the interactions between sub-grid and large
scales. The Leonard and Cross terms in RANS modelling are equal to zero. On the other hand, in LES the Cross and Leonard
terms are approximately equal. Finally, ‘‘the Reynolds term’’, Rij , represents the interactions between sub-grid scales.
The decomposition of the SGS stress tensor, as presented above in Eq. (69) is known as Leonard or triple decomposition
[186]. According to Sagaut [27], another way to decompose Eq. (69) is by double decomposition but with a sharp cut off filter.
Hence Eq. (69) can be written as:

sdij ¼ ui uj  ui uj ¼ ui u0j þ u0i uj þ u0i u0j : ð70Þ
|fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} |{z}
C ij Rij

It is necessary to find a satisfactory model for the sum of Cross and Reynolds terms which constitute a significant part of
the turbulence energy spectrum. The accuracy of LES is dependent on the model for the SGS stresses, which should ensure
the accurate transfer of energy between unresolved and resolved turbulent scales [32]. The suitability of an ideal sub-grid
scale model depends on several properties given by Boris [189] and not repeated here.
In the literature, there are many available SGS models for treating the Reynolds and Cross terms, but none of them can
provide a fully satisfactory solution for the SGS modelling. The most common selection for SGS modelling is based on the
eddy-viscosity hypothesis. In this category of models, the length and velocity scales are specified and combined with a Bous-
sinesq relationship. The most well known and oldest model of this category is the Smagorinsky model [178], which suffers
from many weaknesses, such as: the failure to predict the inverse energy transfer (backscatter), the tuning of Smagorinsky
constant (Cs), the failure to eliminate the eddy viscosity near to the walls and the requirement of a damping function.
An implementation of the Smagorinsky model along with ‘‘a wall-sensitive’’ length scale for predicting turbulent flow
study past a rotating cylinder has recently been investigated by Karabelas [190]. In Fig. 7, the selected O-type grid for the
simulations is presented together with the system of reference for the numerical set up.
In that Karabelas’ study, the physics and the load performance of a rotating cylinder subjected to uniform flow are
explored. The flow is resolved by LES (Fig. 8) and the Re number based on the diameter is equal to 150,000. Based on this
Re number, the flow is characterised as ‘‘separated turbulent flow’’ implying that it becomes turbulent close to the point
of separation. Further downstream the wake is fully turbulent. One of the major outcomes from that study is that the rota-
tion of the cylinder alters the place and nature of transition. At high rotational rates, turbulence is generated without being
necessarily accompanied by immediate separation of the flow. The total resolved kinetic energy is plotted in Fig. 9. The high
values of kinetic energy in the wake are not only due to turbulence but also due to the considerable vortex shedding. Phase
averaging is needed to distinguish the high frequency from the low frequency modes of the fluid motion.

Fig. 7. Computational domain (side section) and the system of reference for the numerical simulations. The azimuth angle h is measured in a clockwise way
while the cylinder rotates in the counter-clockwise direction. A panoramic and a close-up view of the structured grid adopted are also presented.
Reproduced with the author’s permission [190].
712 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Fig. 8. Time-averaged dimensionless stream-wise velocity along the centerline y = 0. Comparison of the numerical results against the Breuer’s and
experimental data of Cantwell and Coles, reproduced with the author’s permission [190].

Fig. 9. Illustration of the total kinetic energy of the fluctuations kf plotted in the near wake for the examined spin ratios, reproduced with the author’s
permission [190].

In order to cope with the limitations of the classical Smagorinsky model an improved dynamic version has been devel-
oped by Germano et al. [191]. The main idea behind the dynamic version is that the coefficients of the model are specified as
the calculation progresses, such that a suitable local value for Cs is determined. The aforementioned idea is accomplished by
a second filter operation (test filter) to the already filtered equations of LES. The main advantage of this procedure is the good
behaviour near to the walls (eddy viscosity tends asymptotically to zero). However, this approach leads to a system of equa-
tions which, in general, is very difficult to satisfy with a single constant Cs. The selected treatment of the aforementioned
deficiency distinguishes the newer versions of the model. The Lilly model [192] adopts a least-square approach, in order
to minimize the error for constant Cs and is the most widely used approach. However, both versions of the model may suffer
from large fluctuations in Cs, which can create instabilities and as a result a local averaging in homogeneous directions is
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 713

necessary. In order to avoid it, the use of the localised dynamic model by Ghosal et al. [193] or the Lagrangian dynamic model
by Meneveau et al. [194] is recommended.
Another important class of SGS models accounting for the backscatter phenomenon is the similarity models
[188,195,196]. This category of models assume that the estimation of SGS tensor can be achieved by using a second filtering
technique with filter size equal to or larger than the initial filter for the resolved scales. The most well known model of this
class is the Bardina model [195] using polynomial inversion [197]. The Bardina model and, in general, all the similarity mod-
els are characterised by the presence of the backscatter phenomenon. However, the similarity models under-predict the SGS
dissipation and demand high computational cost because of the several explicit filtering processes involved.
An alternative class of models that combine the good characteristics of eddy-viscosity and similarity models, is known as
mixed models [198–200]. Thus, a mixed model is obtained by introducing an additional expression for eddy-viscosity in a
similarity model. Mixed models exhibit better results compared to the classical Smagorinsky model and also present the
backscatter phenomenon. Mixed models have been proposed by Zang et al. [198] and Vreman et al. [199], among others.
Another type of SGS models which is based on the high-order approximation of the inverse filter is known as deconvo-
lution models [201,202]. The Approximate Deconvolution Model (ADM) by Stoltz and Adams [201], in which a high-pass fil-
ter is adopted for modelling the interaction between the unresolved and resolved scales, has been tested for channel flows
[201], incompressible wall-bounded flows [203] and compressible flows [202]. The numerical results are compared to DNS
data and exhibit good agreement. More applications of the model may also be found in the work of Adams et al. [204] and in
the textbook by Geurts [205].
Three more classes of SGS models are the Structure function models [206,207], the SGS velocity models [208,209] and the
regularisation models [210–215]. One of the advantages of the latter class of model is that the regularisation principle [213]
allows a transparent modelling between the model equations of the system and the Navier–Stokes equations. As a result,
fundamental properties of N–S equations can be shared between them. In addition, there is no need for any model coefficient
or width of test filter. In this category, the most widely used model is the Leray-a [211,216] which presents a general reg-
ularisation form of N–S equations. It is also worth mentioning that Leray-a model has been used as a closure model for large
Reynolds number channel and pipe flows [212–214]. The numerical results obtained appear in excellent agreement with
empirical data from large Reynolds number flows [217].
Wall resolving LES is a method that near to the wall, where the energy spectrum is made up of anisotropic currents, per-
forms a direct numerical simulation; away from the wall isotropic currents make up a large fraction of the energy spectrum
and are modelled by the use of SGS model. Therefore, the mesh pitch away from the walls may be greater and computational
time is shorter than required by direct numerical simulation. Hence, LES is only meaningful on a refined grid; on a regular
grid fine enough to resolve wall currents the subgrid scale model is unnecessary.
The required resolution of a wall-bounded resolved LES should be sufficient enough in order to resolve the wall layer
according to Piomelli [218]. More specifically, Dx+ 50–150, Dz+ 15–40 and the first point in the wall normal direction
is at y+ < 1. However, with the help of the appropriate wall model the required resolution can be adjusted to Dx+ 100–
160, Dz+ 100–300 and the first point in the wall-normal direction should be at y+ = 30–150 [218]. More details for wall
modelling may be found in the review paper by Piomelli and Balaras [219].

2.6.1. Validation of the LES approach


One of the most important issues of LES approach is the validation of the model, mainly for accuracy, sensitivity and effi-
ciency. A common approach for validation is based on the capability of a given LES to present satisfactory results on a class of
flows [27].
Two related issues are the efficiency and sensitivity of LES approach. Sensitivity expresses the robustness of the method,
which is the variability of the results according to model parameters such as the mesh size, explicit constants, etc. Efficiency
estimates the cost that should be available in order to achieve a given level of accuracy for the various flows.
Accordingly, the above-mentioned issues denote that a good LES depends strongly on the definition of the error. There-
fore, a LES approach can be very good for one given error measure and bad for another. For instance, a LES method can pro-
duce satisfactory results for engineering applications and very bad results for high-level statistics. As a result, the validation
methodology should be explicitly associated with a clear purpose in terms of future use. The validation process for LES
approach is characterised by two alternatives [26]:
A priori validation: In this approach data from a DNS database are used to give exact solutions, in order to compare with
various hypotheses or models in a purely static way. Hence, this method presents good results relating to the nature of non-
linear interactions, but is characterised by the disability to predict the time properties of sub-grid closures.
The second alternative is the a posteriori validation. This approach is characterised by performing numerical simulations
with LES method and comparing its results with a reference solution. This dynamic approach takes all the numerical factors
into consideration.
Geurts and Leonard [220], for reasons described in their paper, proposed the following three criteria for a sound and pre-
dictable LES: (a) adjusting the filter width while adjusting the grid resolution; (b) avoid numerical methods with dissipative
characteristics; (c) implementation of the dynamic modelling approach.
The full validation problem for a given flow is equivalent to finding the space of the solutions spanned by a given LES
method and calculating the solution(s) with the minimum error [27]. The control and estimation of the error in LES which
is generated by the SGS modeling and the adopted numerical method (numerical errors) are difficult to deal with. A recent
714 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

attempt to quantifying the error by the sub-grid modelling and discretization method was presented by Geurts [221,222]. In
addition, Celik et al. [223] proposed the LES_IQ method which represents an index of LES quality. Recently, several attempts
for the quantification of the error in LES have been undertaken by Geurts and Frohlich [224], Meyers et al. [225] and Freitag
and Klein [226], who described inroads into attributing errors to numerical methods versus SGS models [227].

2.7. Monotone Integrated LES (MILES) and Implicit LES (ILES)

The implementation of filtering operation can be invoked explicitly and implicitly. In explicit LES an extra forcing term
(SGS model) is added to the Navier–Stokes equations in order to eliminate the resolution error. On the other hand, in Implicit
LES (ILES) no extra term is added to the Navier–Stokes equations, but the numerical method is selected in such a way that the
numerical and resolution errors will cancel out [26].
Instead of explicit SGS model a class of Non-oscillatory Finite Volume (NFV) methods is adopted for incorporating the
effects of unresolved scales. These methods preserve the monotonicity to the integral form of N–S equations and, as a result,
the correct rate of energy dissipation on the resolved scales is obtained. In this category of NFV schemes belong the Flux-
Corrected Transport (FCT) by Boris and Book [228], the Monotonic Upstream-Centered Scheme for Conservation Laws (MUS-
CL) by Bram Van Leer [229], the Piecewise Parabolic Method (PPM) by Colella and Woodward [230], the Total Variation
Diminishing (TVD) by Harten [231] and the second-order Godunov method by Colella [232]. The above mentioned schemes,
initially used for predicting compressible flow with high accuracy, have been extended to deal with subsonic compressible
and reactive flows. It is also worth noting the implementation and testing of ILES for incompressible flows [233] (e.g. turbu-
lent channel flow, flow over a cylinder and sudden expansion), supersonic jet flows [35] and large-scale urban flows [234].
Boris [235] introduced the use of the term Monotone Integrated Large Eddy Simulation (MILES) [236]. The basic idea
behind the MILES approach is the use of monotone advection schemes which ‘mimic’ the action of what turbulence model-
ling is supposed to do. Monotone advection schemes which satisfy physical properties (e.g. positivity and causality) preserve
stability regardless of turbulence and can have a minimal LES filter and a built-in sub-grid turbulence model coupled con-
tinuously to the grid-scale errors [236]. According to Boris [236] the aforementioned physical properties are adequate to pro-
vide efficient transfer of the sub-grid motions, by minimizing the effect of the numerical filter of the well-resolved scales in
the resolved grid.
The MILES approach has been adopted for predicting turbulent flows such as beam-channel interactions, open-air ammu-
nition-destruction, shock waves, geophysical and atmospheric flows [236]. It is concluded, after many years of research on
MILES, that the monotone (positivity-preserving) algorithms appear to lead to good agreement with theory and data, being
free of any eddy viscosity or explicit sub-grid model. MILES is very useful for complex problems which include compressible
and multi-physics flows with complex geometries because it is easy to implement.

2.8. Unsteady Reynolds-Averaged Navier–Stokes (URANS)

Unsteady Reynolds-Averaged Navier–Stokes (URANS) or Transient Unsteady Reynolds-Averaged Navier–Stokes (TRANS)


is an alternative approach to LES. URANS is characterised as the most complex version of RANS. In URANS, a time solution of
the conventional RANS is performed for 3D unsteady problems, with or without special treatment for the flow instability
[85]. The equations of URANS are similar to the RANS models but include the additional unsteady term. The most important
aspect for an URANS model is the correct selection of the suitable turbulence model and the discretization scheme (which
must be not too dissipative). URANS performed well for vortex shedding behind bluff bodies [237] and surface mounted cube
[238] and around cars [234], among other applications. Application of TRANS for predicting environmental turbulent flows
and pollutant dispersion have been investigated by Hanjalic and Kenjeres [160] with good results. For more details, the inter-
est reader is directed to the reviews by Speziale [240], Spalart [241], and the recent one by Kenjeres and Hanjalic [242].
The second generation URANS (2G-URANS) models, according to the terminology of Frohlich and von Terzi [36], is also
worth mentioning. They are the Scale-Adaptive Simulation (SAS) model [243] and the Partially Filtered Navier–Stokes
(PANS) model [244]. The former is based on revisiting the k–kl model of Rotta [245] by Menter and Egorov [246], and the
latter is based on the transformation of a RANS model into a URANS by just changing the dependant coefficients.
The PANS model is now well proven with many practical and complex turbulent flows applications such as channel flow,
external car aerodynamics, and simplified train geometry with a crosswind [247]. Recently, a low-Re number version of the
model has been developed by Ma et al. [248] and another version with an embedded LES model has been proposed by David-
son and Peng [249].
The SAS model was modified by Menter and Egorov [246] with the use of the k–x SST model and as a result a new version
of the model was developed, known as SST-SAS. The model has been tested for different type of flows such as axisymmetric
hill, channel and asymmetric diffuser [250]. It is concluded that the SST-SAS is an equivalent VLES (introduced next in Sec-
tion 2.9), because it represents the spectrum of resolved scales between LES and URANS [175].
A similar model to PANS is the Partially Integrated Transport Model (PITM) [251,252] with both models using RANS and
LES framework. In the LES framework both models avoid to adopting a length scale as the width of a filter function. These
models resolve a large part of the turbulence energy spectrum and as a result they are named as second generation URANS
[36]. The difference of PANS against PITM is that in the first the coefficients which are responsible for the turbulence diffu-
sion in k and e are modified. The PITM model has been validated against numerical and experimental data for the decay of
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 715

homogeneous turbulence, fully turbulent channel flow and unsteady turbulent flow with periodic forcing, with promising
and encouraging results [251,252].

2.9. Very LES (VLES) and Detached-Eddy Simulation (DES)

The Very Large Eddy Simulation (VLES) was proposed by Speziale [240], and was refined and reformulated by Johansen
[253]. VLES combines RANS and LES in such a way that use is made of the advantages of those approaches. The combination
of these strategies involves numerical resolution in time and space only for the very large eddies, while a significant part of
the turbulence spectrum requires modelling [85]. VLES constitutes a turbulence approach with coarser grid requirements
compared to LES and the idea of VLES originates from the atmospheric-science community. The main advantage of this
method is that it can preserve efficiency and accuracy within reasonable computer-time demands.
A similar approach with the same characteristics of VLES is the Partially Resolved Numerical Simulation (PRNS) technique
by Liu and Shih [254]. PRNS is based on the assumption that the small scales of turbulence have small timescales and as a
result a filter with a fixed width is adopted in order to distinguish the large scales of turbulence. Hence, according to the
value of the temporal filter width the PRNS approach operates between RANS and LES modes. The PRNS approach was tested
for turbulent pipe flow [175,254], non-reacting flow in a single injector flame tube [254], internal reacting and external static
stall flows [255].
There are more approaches that follow the VLES methodology but it is not possible to describe all of them, due to space
limitations. However, the interested reader may find more information in the Limited Numerical Scales (LNS) by Batten et al.
[256], the approach by Ruprecht et al. [257], the self-adapting turbulence model by Perot and Gadebusch [258,259] , the
method by Hsieh et al. [260], and the recent turbulence modelling approaches by Labois and Lakehal [261] and Han and Kraj-
novic [262].
The Detached-Eddy Simulation (DES) originated in 1997 [263] and was first used in 1999 [264], with the purpose of cop-
ing with massively separated flows at high-Reynolds number. DES combines LES and RANS approaches, based on the turbu-
lence length scale and the grid spacing. Hence, LES is used for regions of massive separations and RANS within the boundary
layer. The official definition of DES, in accordance with Travin et al. [265] is ‘‘A three-dimensional unsteady numerical solu-
tion using a single turbulence model, which functions as a sub-grid scale model in regions where the grid density is fine
enough for a large-eddy simulation and as a Reynolds-averaged model in regions where it is not’’. More details about the
equations, advantages, limitations and implementation of DES can be found in the review paper of Spalart [264].
Improvements of the DES model such as Delayed DES (DDES), Improved DDES (IDDES) and SST-DDES have been presented
for different case studies. DDES is characterised as a modified version of DES approach in order to tackle the nonphysical
behaviour in the vicinity of the boundary layers. Therefore, a more generic formulation of the ‘‘shielding’’ function was pro-
posed which is dependent only on the wall distance and the eddy viscosity. The term IDDES was originated by Shur et al.
[266], who indicated that the original DES could be transformed with wall modelled LES capabilities. Finally, the SST-DDES
constitutes the combination of DDES by Spalart [267] with the SST-DES approach of Strelets [268]. More details, applications
(e.g. sharp edged delta wings, bluff bodies, ground vehicles, active flow control by suction/blowing, vibrating cylinders with
strakes, cavitation inkjets, building, air inlets, aircraft in spin, high-lift devices [264]) and modifications for the above men-
tioned versions of DES can be found in the works of Spalart et al. [267], Spalart [264] and Gritskevich et al. [269].

2.10. Hybrid RANS/LES strategies

All the aforementioned approaches in the last two Sections 2.8 and 2.9 are expressed by the combination of LES with
RANS models. According to Leschziner et al. [270] all those approaches, excluding very few cases, can be divided into four
basic types; wall laws or wall functions, zonal schemes, seamless schemes and hybrid schemes.
The hybrid schemes were introduced in order to overcome one of the main disadvantages of LES for wall-bounded tur-
bulent flows. For a wall-bounded turbulent flow the demand of fine grid resolution is great, in particular for high Reynolds
number flows, in the vicinity of the walls. Therefore, the hybrid RANS/LES strategies decrease the demand for fine grid res-
olution at the wall region. This can be achieved by using a low-Re RANS model at the wall region, while for the outer region
an LES approach is adopted. Thus the near wall turbulence is modelled instead of being fully resolved and thus a coarser grid
may be used.
Much work has been dedicated to Hybrid RANS/LES approaches [271–276]. Due to space limitation the interested reader
is directed to the review paper by Frohlich and von Terzi [36],

3. Applications of DNS and LES to flows in pipes and flows with a free surface

This section is devoted to listing interesting applications of the DNS and LES approaches to two well-researched flow
types. This discussion is valuable (a) in order to present interesting practical applications of these two important turbulence
modelling methods, (b) in order to highlight some important turbulent-flow physics, and (c) in order to present several more
references on turbulence modelling applications, thus completing the present review. LES and particularly DNS applications
716 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Table 2
Previous studies in DNS of turbulent single-phase pipe flows.

Reference Mesh Reb Geometry Method


Nikitin [279] 43  64  43 2000–4000 Straight pipe FDM + SM
Eggels et al. [280] 96  128  256 5300 Straight pipe FVM
Zhang et al. [281] 75  128  128 2500, 4000 Straight pipe SM
Loulou [282] 72  160  192 5600 Straight pipe B-spline/spectra
Orlandi and Fatica [284] 128  96  257 4900 Rotating pipe FDM
Orlandi and Ebstein [285] 129  96  193 4900 Rotating pipe FDM
Schmidt et al. [286] Cyl: 150 elements 4910 Straight pipe SEM
Car: 64 elements
Wagner et al. [288] 70  240  486 10,300 Straight pipe FVM
Fukagata and Kasagi [289] 96  128  256 5300 Straight pipe FDM
Veenman [290] 152  256  394 10,300 Straight pipe SM
Nikitin and Yakhot [291] 200  160  256 6000 Elliptical pipe FDM
Voronova and Nikitin [293] 32  256  64 4000 Elliptical pipe FDM
Voronova and Nikitin [294] 48  256  128 6000 Elliptical pipe FDM
Wu and Moin [295] 300  1024  2048 44,000 Straight pipe FDM
Boersma [296] 430  512  1024 24,500, 61,000 Straight pipe SM + FDM
Wu et al. [297] 256  1024  2048 24,580 Straight pipe FDM
Khoury et al. [298]a 18.67  106, 121.4  106, 4374  106, 2.184  109 5300, 11,700, 19,000, 37,700 Straight pipe SEM

FDM: Finite Difference Method.


FVM: Finite Volume Method.
SM: Spectral Method.
SEM: Spectral Element Method.
Car: for Cartesian coordinates.
Cyl: for Cylindrical coordinates.
Reb: Reynolds number based on bulk-mean velocity and pipe diameter of the pipe.
a
Number of grid points for each considered Reb, respectively.

lead to much more detailed results than those obtained by RANS, as they allow the computation of high order statistics and
they also reveal secondary flows that are very difficult for RANS model to reveal.

3.1. DNS of turbulent pipe flows

Turbulent pipe-flow numerical studies (Table 2) historically have not been as popular as channel flows. This is due to the
strong restrictions imposed on the time step choice and the very fine grids required to simulate pipe flows, in order to elab-
orate on the singularity at the centerline. Some available turbulence pipe experimental studies can be found in the review
papers by Marusic et al. [277] and Mullin [278].
An early attempt to perform DNS computations of three-dimensional turbulent flows in pipes was performed by Nikitin
[279]. In that work, a mixed finite differences and spectral method was used for the numerical simulation of the flow within
a range of Reynolds number from 2000 to 10,000. The numerical results, which describe the evolution of motion and the
characteristics of the flow, presented fair agreement with the available experimental data.
The same year, Eggels et al. [280] presented DNS of fully turbulent pipe flow at3 Reb = 5300 in a computational domain
with length of 10R. Statistical results on DNS were investigated in order to examine whether the axisymmetric geometry affects
the velocity fluctuations of the flow or not. The numerical results were compared with the DNS data of channel flow by Kim
et al. [58] at the same Re number, with the aim to observe any possible differences between axisymmetric pipe and the plane
channel geometries. The DNS results were also validated with experimental data obtained by PIV and LDA measurements. It was
noticed that the statistics of velocity fluctuations exhibit less effects for the pipe compared to plane channel geometry. However,
it was observed that the turbulence-intensity differences appear to be negligible between the pipe and channel flow. The skew-
ness factor differs significantly between the two geometries, caused probably by the impingement mechanism of the wall due to
the transverse curvature effect. The high order statistics data and the energy budget computations between the channel and
pipe geometries exhibit fair and excellent agreement with available experimental data, respectively.
DNS for fully turbulent pipe flows has also been investigated by Zhang et al. [281] and by Loulou [282]. The former used a
full spectra method and presented features of turbulent pipe flow at Re = 2500 and 4000, while the latter used a B-spline/
Fourier embedded method with the divergence-free Galerkin method of Leonard and Wray [283] at Reb = 5600.
Orlandi and Fatica [284] performed DNS using finite difference methods for investigating the effects of flow in a rotating
circular pipe. The rotation number, N, was in a different range compared to the work of Eggels [280], but was not high
enough to investigate the near-wall vortex structures. The numerical results in the non-rotating case were validated by
the data of Eggels [280]. They concluded that a strong interaction exists between numerical and sub-grid dissipation. How-
ever, a degree of drag reduction demonstrated by numerical and experimental data and the differences of turbulent statistics
were explained by the tilting of the near-wall -vortex -structures in the direction of rotation.

3
Reynolds number based on bulk velocity and pipe diameter.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 717

Fig. 10. Visualization of the turbulent pipe flow over the surface of 1  r = 0.1 using contours of instantaneous uz. Red represents higher values of uz.
(a) Reb = 5300. (b) Reb = 44 000 [295]. X. Wu, P. Moin, A direct numerical simulation study on the mean velocity characteristics in turbulent pipe flow,
Journal of Fluid Mechanics 608 (2008) 81–112, reproduced with permission. (For interpretation to colours in this figure, the reader is referred to the web
version of this paper.)

Extension of the previous study was the work of Orlandi and Ebstein [285] with the increase of the rotation number, N, up
to 10 and the derivation of Reynolds stress budgets. No significant changes compared to previous results were observed
[284].
Schmidt et al. [286] investigated a DNS of turbulent pipe flow at Reb = 4910, by employing a spectral element method, in
both cylindrical and Cartesian coordinates. The numerical results were validated by available experimental and numerical
data of Den Toonder and Nieuwstadt [287] and Eggels et al. [280], respectively. The mean velocity profile presented excellent
agreement against the experimental data by both methods, while urms values are almost identical for both cases, excluding
some small differences at the outer layer. Finally, the shear stress distribution did not exhibit almost any differences among
all data.
Wagner et al. [288] conducted DNS for fully turbulent pipe flow at Reb = 10,300, by using a finite volume technique, sim-
ilar to Eggels et al. [280]. Turbulence quantities were investigated and showed significant low Re effects, particularly for the
turbulence kinetic energy budget, the Reynolds stress tensor, vorticity and pressure fluctuations and the two point correla-
tions for velocity.
The same year, Fukagata and Kasagi [289] presented DNS results at Res = 180 by using a second-order finite difference
method. The length of computational domain was 10R and the adopted mesh was 96  128  256 with Dr+ = 0.46 (wall)
to Dr+ = 2.99 (center). Their numerical results compared with DNS data by Eggels et al. [280] and DNS data of channel flow
by Moser et al. [59], presenting good agreement.
Some years later, Veenman [290] investigated the DNS of turbulent pipe flow at Reb = 10,300 by using a pseudo-spectral
approach. The numerical results obtained focused on statistical parameters which are important for the development of
Lagrangian stochastic turbulence models. The proposed stochastic turbulence model was used to describe the dispersion
flow of a passive scalar from a point source in a pipe. The obtained dispersion statistics showed satisfactory agreement with
available DNS and experimental data.
Nikitin and Yakhot [291] presented DNS of turbulent flow in elliptical ducts. They implemented Immersed Boundary
Method (IBM) according to the research of Kim et al. [292] and estimated that the mean flow characteristics, the Reynolds
718 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Fig. 11. Pseudo-colour visualisation of the instantaneous axial velocity uz normalised by the bulk velocity Ub. (a) Reb = 5,300; (b) Reb = 11,700;
(c) Reb = 19,000; (d) Reb = 37,700. Here, the colours vary from 0 (black) to 1.3 (white) [298]. ‘‘G.K. Khoury, P. Schlatter, A. Noorani, P. Fisher, G. Brethouwer, A.
Johansson, Direct numerical simulation of turbulent pipe flow at moderate high Reynolds numbers, Flow, Turbulence and Combustion 91 (2013) 475–495.
This is Fig. 4 in the publication in which the material was originally published. With kind permission from Springer Science and Business Media.’’
(For interpretation to colours in this figure, the reader is referred to the web version of this paper.)

stresses and turbulence intensities present similarities compared to data from channel flow. In addition, a reduction of tur-
bulence effect near to the wall region of the major axis was noticed due to the transverse curvature effect. Moreover, the
presence of instantaneous velocity structures near the wall region was identified as streaks. However, these structures were
not present near to the major axis endpoints.
Extension of the previous study was the work of Voronova and Nikitin [293,294]. Both studies performed DNS in elliptical
ducts but for higher Reynolds number, 4000 and 6000, respectively.
Wu and Moin [295] performed DNS (Fig. 10) for fully turbulent pipe flow at Reb = 5300 and 44,000 based on bulk velocity,
by using a finite difference technique of second order on 630  106 grid nodes. A computational domain with length equal to
15R was used at both Re numbers, while the mesh at Reb = 5300 and 44,000 was 256  512  512 and 300  1024  2048
along the r, h and z directions, respectively. Their numerical results for mean flow statistics appear in good agreement with
available experimental data from Princeton Superpipe at Reb = 41,727 and 74,000, while for the second-order statistics good
agreement is observed at Reb = 74,000.
Recently, Boersma [296] presented DNS results of turbulent pipe flow at Reb = 24,500 and 61,000, by using a combination
of a pseudo-spectra method and a 6th order finite difference method. The pseudo-spectra method was used for the axial and
azimuthal directions while a 6th order finite-difference method was adopted for the radial direction. Turbulence statistics
were computed and autocorrelation functions and one-dimensional energy spectra were also presented. The numerical
results for the case of Reb = 24,500 appear in excellent agreement with experimental data by Den Toonder and Nieuwstadt
[287].
Continuation of the previous study of Wu and Moin [295] was the work of Wu et al. [297] by using the same numerical
method. They presented numerical results for DNS of pipe flow with a computational domain equal to 30R at Reb = 24,580. A
grid with resolution of 256  1024  2048 was used along the r, h and z directions, respectively. The numerical results high-
lighted the importance of large and very large scale motions. The very large scales indicate streamwise acceleration of the
flow in the vicinity of the wall, based on force spectra. The small scales seem to decelerate the mean streamwise profile
according to observations in force spectra of previous experimental studies. It was the first time that net force spectra com-
putations were performed in the buffer layer.
Khoury et al. [298] performed DNS for investigating incompressible flow in a smooth pipe of radius R and length 25 R, by
using a high-order spectral element method. Numerical results were obtained at Reb = 5300, 11,700, 19,000 and 37,700. In
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 719

Fig. 12. (a) Pseudo-colours of the instantaneous axial vorticity xz for Reb = 19,000 together with the spectral element boundaries. (b) Zoomed view of an
upper right part of (a) [298]. ‘‘G.K. Khoury, P. Schlatter, A. Noorani, P. Fisher, G. Brethouwer, A. Johansson, Direct numerical simulation of turbulent pipe flow
at moderate high Reynolds numbers, Flow, Turbulence and Combustion 91 (2013) 475–495. This is Fig. 5 in the publication in which the material was
originally published. With kind permission from Springer Science and Business Media.’’ (For interpretation to colours in this figure, the reader is referred to
the web version of this paper.)

Fig. 11, the instantaneous axial velocity is presented, according to the increase of Re number and therefore the increase in the
range of the scales. Fig. 12 depicts the instantaneous axial vorticity at Reb = 19,000 with the numerical grid. Strong vortex
motion in the vicinity of the walls is observed as well as the configuration of small rotating vortices. The numerical results
were extensively compared with numerical data from pipe, channel and boundary layer turbulent flows. Turbulence statis-
tics such as turbulence kinetic energy budgets and velocity fluctuations were also presented and evaluated. It was concluded
that the variation of pressure and velocity fluctuations depend on the Reynolds number for the examined flows. High degree
of similarity was exhibited for all the considered types of flows and important differences were observed in the wake region.
Turbulence kinetic energy budgets present independence, in the inner region, from the type of flow up to y+ 100, along
with great differences for the wake region.

3.2. DNS of turbulent free-surface flows

Some important DNS work involving turbulent flows with free surfaces, with and without shear is listed here.
Lam and Banerjee [299] conducted DNS with the use of a Fourier–Chebyshev pseudo-spectral method for the examination
of the influence of shear, and of boundary conditions, on the turbulent structures near to the wall and the free slip surface.
The aim of their research was focused on the determination of the critical parameter which was responsible for the config-
uration of turbulent streaky structures. However, they did not examine the relationship between turbulence structure and
scalar transport mechanism.
Komori et al. [300] performed DNS calculation for the clarification of turbulence structure at the interface of an open
channel flow. In contrast to the study of Lam and Banerjee [299] the proposed methodology describes the mechanism of sca-
lar transfer for a gas–liquid interface with zero-shear. The numerical simulations were conducted with the use of finite dif-
ference method and a Boundary Fit Coordinate (BFC) system without any approximations. The numerical results for the
bursting frequency and mass transfer coefficient were compared with turbulence statistics by Laser Doppler Velocimetry
(LDV) and author’s previous studies, presenting good agreement. They also noticed that in the vicinity of the wall there were
large eddies. These large eddies are lifted up towards the interface position and renew it. Furthermore, they promote the
mass transfer across the gas–liquid interface.
Lombardi et al. [301] investigated via DNS the coupling between gas and liquid phase, with keeping flat the interfacial
region (Fig. 13) of the flow. The numerical results focused on the turbulence statistics and the clarification of the mechanics
of coupling flows between the phases. It was concluded that the turbulence structure on the gas side has great similarities
with that at the wall. They also observed that in some applications of two-phase flows the gas side might perceive the liquid
phase as a solid surface.
De Angelis [302] presented a continuation of the research of Lombardi et al. [301] with the aim at extending his work to a
non-flat interfacial region, while the stratified flow was accounted for with a freely deformable interface.
De Angelis et al. [303] also examined stratified flows with the use of wind stresses on the gas–liquid interfacial region,
while the wavy interface was transformed to a simple geometry in order to facilitate a simple computation. They concluded
that the scalar exchange rates depended in a similar way on the activities on the gas side, with high exchange rates occurring
with ‘‘sweeps’’, whereas low exchange rates with ejections [304].
Fulgosi et al. [305] performed DNS to investigate the turbulence near a freely deformable interface (Fig. 14). The research
was focused on the interfacial sub-layer and particularly on the gas side, with the aim at identifying how the wavy-induced
mechanisms affect the flow characteristics. Numerical results on the gas side were compared with channel flow data at the
720 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Fig. 13. Sketch of the physical problem [301]. Reprinted with permission from P. Lombardi, V.D. Angelis, S. Banerjee, Direct numerical simulation of near-
interface turbulence in coupled gas-liquid flow, Physics of Fluids 8 (1996) 1643–1665. Copyright [1996], AIP Publishing LLC.

Fig. 14. Sketch of the simulated problem. The elevation of the waves has been amplified by a factor 5 [305]. M. Fulgosi, D. Lakehal, S. Banerjee, V.D. Angelis,
Direct numerical simulation of turbulence in a sheared air-water flow with a deformable interface, Journal of Fluid Mechanics 482 (2003) 319–345,
reproduced with permission.

same shear Reynolds number. However, the numerical method was not the most appropriate for flows with large deformable
interface, due to the possible entrainment of the liquid phase and the formation of large waves. It was observed that the
numerical results do not exhibit great differences compared to the previous study of Lombardi et al. [301].
Banerjee et al. [306] presented DNS for the clarification of the transfer mechanisms at deformable interfaces. They
increased the friction velocities with the purpose to present surface deformations of high waveslops, up to the point which
they do not lead to wave breaking. In the cases of high shear rate because of gas, the turbulence is presented in the vicinity of
the interface, as being similar to that at solid boundaries. Fig. 15 illustrates contours of instantaneous interfacial heat transfer
coefficient for wavy and flat interface. The top cases are for the gas phase and the other two at the bottom for the liquid
phase. It was concluded that the surface divergence model is suitable for the prediction of gas transfer for non-sheared inter-
face, but also for wind-shear cases.
Lin et al. [307] developed an air–water coupled method in order to simulate the interaction of two fully developed
turbulent layers (air and water), above and below the interface. In that study the wind-wave generation processes were
investigated by performing DNS and coupling of two turbulent flows (air and water) across a deformable interface. Limita-
tion of the model is the linearization of the boundary condition at the interface, so that the model is suitable only for
small-amplitude waves. It was observed that in the initial stage of the simulated waves there were great similarities of their
characteristics with field and experimental data.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 721

Fig. 15. Instantaneous patterns of the interfacial heat transfer coefficient at the flat and wavy interface. (upper panel) gas side, (lower panel) liquid side
[306].

Numerical simulations of turbulent condensing vapour–liquid flow were conducted by Lakehal et al. [308]. A DNS
approach was selected for the investigation of turbulence characteristics of the two-phase flow, accounting for condensation
effects. The numerical results showed that the effects of condensation are important on the turbulence characteristics of the
flow for both phases (liquid and vapour). More specifically, the turbulence statistics and Reynolds stresses increase at the
vapour side, while they decrease at the liquid side. Fig. 16 depicts the coherent structures with and without condensation
for both phases of the flow. The results for the obtained velocity profiles were compared with DNS and experimental data.
It was also observed that condensation affects the configuration of interfacial waves by weakening them.
Trontin et al. [309] performed DNS of two-phase interfacial flows. The main goals of their study were to investigate the
interaction between small turbulence scales and the deformable interface. The pioneering part of the research was the pre-
sentation for the first time of DNS results for the interface/turbulence interactions, in case of a widely deformed interface. A
energy budget and turbulence statistics were also exhibited.

3.3. LES of turbulent pipe flows

LES of three-dimensional turbulent pipe flow studies (Table 3) are not as numerous as for channel flows due to the for-
mer’s centreline singularity.
An early attempt to perform LES of turbulent pipe flow was undertaken by Unger and Friedrich [310] for complex geom-
etries at Reb = 50,000, by using second-order finite volume method. The classical Smagorinsky model was used on a mesh
size of 96  128  256. Numerical results for instantaneous turbulence quantities were presented.
Eggels and Nieuwstadt [311] implemented a Smagorinsky SGS model for a LES approach and showed that their numerical
results were in good agreement with experimental data.
Boersma and Nieuwstadt [312] investigated the effects of turbulent flow in a curved pipe, by means of LES technique and
the classical Smagorinsky SGS model at Reb = 20,000. Numerical results for turbulence statistics, mean velocity profile and
secondary motion of the flow were presented and discussed; they were also compared with available experimental data,
presenting satisfactory agreement.
722 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Fig. 16. Three-dimensional distribution of vortical structures in the presence of interfacial mass exchange, made visible by the use of isosurfaces of
k2 = 0.03. [(a) and (b)] Case R2. [(c) and (d)] Case C21. The amplitude of the waves has been magnified ten times in the vertical direction [308]. Reprinted
with permission from D. Lakehal, M. Fulgosi, S. Banerjee, G. Yadigaroglu, Turbulence and heat exchange in condensing vapor-liquid flow, Physics of Fluid 20
(2008) 065101–18. Copyright [2008], AIP Publishing LLC.

Table 3
Previous studies in LES of turbulent single-phase pipe flows.

Reference Mesh Reb Geometry SGS model Method


Unger and Friedrich [176] 96  128  256 50,000 Straight pipe Smagorinsky FVM
Eggels and Nieuwstadt [177] 96  128  256 59,500 Rotating pipe Smagorinsky FVM
Boersma and Nieuwstadt [190] 40  114  200 20,000 Curved pipe Smagorinsky FVM
Yang [191] 192  64  128 20,000 Rotating pipe Smagorinsky FVM
Dynamic
Rudman and Blackburn [192] Car: 192 elements 36,700 Straight pipe Smagorinsky SEM
Schmidt et al. [154] Cyl: 80 elements 16,000 Straight pipe Smagorinsky SEM
Car: 105 elements
Feiz et al. [193] 65  39  65 4900, 7400 Rotating pipe Smagorinsky FDM
Dynamic
Jordan [194] 64  141  401 8000 Pipe roughened (p/k = 5 ribs) Dynamic FDM
Vijiapuraru and Cui [195] 64  96  64 5000, 30,000 Straight pipe Smagorinsky FDM
Dynamic
Vijiapuraru and Cui [196] 96  128  128 100,000 Pipe roughened (p/k = 2, 5, 10 ribs) Dynamic FVM
Jung and Chung [197] 128  256  256 7000–36,000 Straight pipe Dynamic FDM

FDM: Finite Difference Method.


FVM: Finite Volume Method.
SEM: Spectral Element Method.
Car: for Cartesian coordinates.
Cyl: for Cylindrical coordinates.
Reb: Reynolds number based on bulk-mean velocity and pipe diameter of the pipe.
p/k: ratio of rib periodicity where p is the distance between two successive ribs and k is the rib height.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 723

Table 4
List of the described numerical approaches for simulating turbulent flows in the present review paper, excluding the cases for pipe flows (DNS and LES) which
are presented in Tables 2 and 3, respectively.

Approaches-models Type of flow


DNS
Lilly [55], Orszag and Patterson [56], Rogallo [57], Kim et al. [58], Moser Two-dimensional turbulence [55], homogeneous isotropic turbulence
et al. [59], Spalart [64], Schlatter et al. [65], Kaneda et al. [72], Kaneda [56,57,72,73], channel [58–62,71], boundary layer [63–69]
and Ishihara [73], Abe et al. [60], Del Alamo et al. [61], Hoyas and
Jimenez [62], Lee et al. [71], Schlatter and Orlu [67], Sillero et al. [68]
RANS
Zero-equation models
Prandtl’s mixing length [88], Cebeci-Smith [89], Baldwin-Lomax [90] Boundary layer [88,89], flat plate and airfoil [90]
Half-equation models
Johnson and King [92–95] Pressure driven separated flows [92], transonic separated flows [93],
compressible [94]
One-equation models
Baldwin and Barth [96,99,100], Spalart and Allmaras [97,98,101,102], Airfoil [99–102], wake, jet, boundary layer [105]
Fares and Schroder [105]
Two-equation models (including LRN)**
Standard k–e [107], realisable k–e [109], RNG k–e [110], LRN k–x Jets, mixing layers, channels and boundary layers [109], pollutant
[108,131], Wilcox (2006) k–x [86], SST [119–121], Lam-Bremhorst dispersion [111–114], street canyon [115], rotating cylinder [116],
k–e [129], Bredberg et al. k–x [130], Pang et al. k–x [132] turbomachinery blades, wind turbines, free shear layers [119–121]
NLEVM
Quadratic models [134,135], Cubic models [137,138] For separated flows in adverse pressure gradient, turbine blade, flow
plate, jets [134–138]
Advanced EVM
Durbin’s u2–f model [139,142], f–f model [140], BL- u2/k [143,145] Rotating cylinder, rotating channel flows, axial rotating pipe and square
duct [141]
DSM
Models for the diffusive term (Dij): [146–148] Street canyon flows [115], impingent jet flows [144,162]
The redistribution term (Uij) is usually modelled by IP model [150]. Flow around a car, axially rotating pipe, 180° turned U bend, backward-
Improvements are: LRR-QI model [150], SSG model [152], CL model facing step [162]
[153] and LT model [154]. EBSDM model [144], Durbin’s DSM [159]
and HTM model [162]
ASM
Pope [163], Gatski and Speziale [134], Jongen and Gatski [164] and Wallin 2-D flows [163], 3-D flows [134,164], channel flows, boundary layer,
and Johansson [165] wings and impinging shock [165]
Two-fluid models
Spalding’s model [166], Malin [167], Markatos and Kotsifakis [168], Shen Free shear flows [167], combustion [168], stratified flows [169], flow and
et al. [169], Yu et al. [170], Young’s model [171], heat transfer of air curtains [170,174], Rayleigh–Taylor mixing [171,172]
2SFK model [172], Lin et al. [173] and Cao et al. [174] and UV disinfection reactor flow [173]
LES
Smagorinsky [178], Lilly [179,192], Deardorff [180], Schumann [183], Isotropic turbulence [181,182], channel flow [180,183–
Kraichnan [181], Chasnov [182], Moin and Kim [184], Piomelli [185], 185,201,204,205,212–214,217], wall-bounded flows [203], compressible
Bardina [195], Zang et al. [198], Vreman et al. [199]. Germano [191], flows [202], rotating cylinder [190]
Ghosal et al. [193], Germano [191], Meneveau et al. [194], Stolz and
Adams [201], Stolz et al. [202], Leray-a [211,217,216], Clark-a [210],
Metais and Lesieur [206], Ducros et al. [207], Domaradzki and Saika
[208], Domaradzki and Loh [209]
MILES-ILES
Boris and Book [228], Boris [235,236], Fureby [233], Patnaik et al. [234], Channel flow, flow over a cylinder [233], supersonic jet flows [35], large-
Colella and Woodward [230], scale urban flow [234]
URANS (including 2G-URANS)
Johansson et al. [237], Person and Davidson [238,239], Speziale [240], Vortex shedding behind bluff bodies [237], surface mounted cube [238],
Kenjeres and Hanjalic [242], Hanjalic and Kenjeres [160], SAS model flow around car [239], environmental turbulent flows and pollutant
[243,246], PANS model [244,247–250], PITM [251,252] dispersion [160]
Homogeneous turbulence, channel flow and unsteady flow with periodic
forcing [251,252]
VLES
Speziale [240], Johansen [253], PRNS [175,254,255], LNS [256], Ruprecht Turbulent pipe flow [175,254], non-reacting flow in a single injector
et al. [257], Parot and Gadebusch [258,259], Hsieh et al. [260], Labois flame tube [254], internal reacting and external static stall flows [255]
and Lakehal [261], Han and Krajnovic [262]

(continued on next page)


724 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

Table 4 (continued)

Approaches-models Type of flow


DES
Spalart et al. [267], Spalart [264], IDDES [266], DDES [266], SST-DES [288], Sharp-edged delta wing, bluff bodies, ground vehicles, active flow control
Gritskevich [269] by suction/blowing, vibrating cylinders with strakes, cavitation in jets,
building, air inlets, aircraft in spin, high-lift devices [264]
Hybrid LES/RANS
Leschziner et al. [270], Davidson [271], Hambla [2003], Temmerman et al. Channel flow [271–275], separated flow [273], plane asymmetric diffuser
[2005], Tucker and Davidson [274], Davidson and Dahlstrom [275], flow [275]
Xiao et al. [276]
⁄⁄
LRN: Low-Reynolds number.

Yang [313] performed LES for the investigation of swirl effects driven by a rotational wall at Reb = 20,000. A finite volume
method was adopted with a computational domain of 4D for the numerical simulations. The performance of two sub-grid
scale models, namely dynamic and classical Smagorinsky, has been investigated. The numerical results appeared to be in fair
agreement with measurements and both sub-grid scale models performed equally well.
Rudman and Blackburn [314] and Schmidt et al. [286] presented results from an LES of turbulent pipe flow with bulk flow
Reb = 36,700, with the classical Smagorinsky model and van Driest type wall damping. The numerical results were compared
with experimental measurements, exhibiting fair agreement.
Feiz et al. [315] performed LES of a rotating pipe at moderate Reb numbers 4900 and 7400. The numerical results were com-
pared with simulations obtained by DNS data [284]. Two different SGS models (Smagorinsky and dynamic) were examined for
the prediction of turbulent flow with and without rotation. The best performance was exhibited by the dynamic model.
The same year, Jordan et al. [316] conducted LES for the investigation of turbulent flow inside a cylindrical ribbed duct
with ratio of rib periodicity equal to p/h = 5 and Reb = 8000. They employed a finite difference technique for the discretization
of Navier–Stokes equations and adopted a dynamic model for the description of SGS stress field. Turbulence statistics such as
streamwise intensity, radial intensity and Reynolds stress were computed and compared with DNS data of Eggels et al. [280]
and ‘‘at-sea’’ measurements (Reb = 4  106). The numerical results and ‘‘at-sea measurements’’ presented essentially equiv-
alent pressure core loss at the center of the circular duct, which proves that the turbulence physics for both Reynolds num-
bers are scale-similar.
Vijiapuraru and Cui [317] investigated the fully developed turbulent flow by means of LES for two different Reynolds
numbers, Reb = 5000 and 30,000. A finite difference method was employed for the discretization of Navier–Stokes equations
in cylindrical coordinates, while the classical Smagorinsky and dynamic model were selected for the subgrid scale modelling.
Mean velocity calculations and computations of high order turbulence statistics were presented and validated with available
numerical and experimental data. It was concluded that the proposed methodology is suitable for handling turbulent pipe
flows at low and moderate Re numbers.
Vijiapuraru and Cui [318] examined also turbulent flows in circular ribbed pipes by using LES, and the finite volume
method. The numerical results obtained for Reb = 100,000 were compared with experimental data and presented fair agree-
ment for the considered p/k = 2, 5 and 10. It was concluded that LES is suitable for predicting the turbulence statistics but
that it needs at least one order more CPU time compared to the classical RANS models.
Jung and Chung [319] undertook numerical simulations for the investigation of accelerated turbulent flow in cylindrical
pipe by means of LES. A second order central difference scheme was employed for the discretization of incompressible
Navier–Stokes equations and the numerical simulations were performed for linearly increasing value of Reb from 7000 to
36,000 due to the acceleration. Statistics for the skin friction coefficient, mean velocity, velocity fluctuations and quadrant
analysis were presented and discussed. That LES study was the first attempt to treat temporal acceleration in a fully-devel-
oped turbulent pipe flow and it was concluded that the turbulence anisotropy increased during the temporal acceleration.

3.4. LES of turbulent free-surface flows

To the authors’ best knowledge two-phase pipe LES studies are not available except for the recent work of Lakehal [320].
Lakehal [320] performed Large Eddy Interface Simulation (LEIS) for modelling the slug formation of a two-phase pipe flow. A
BFC grid was adopted for the computational domain and a level set method for the interfacial region. The numerical results
were compared with analytical and experimental data, predicting the slug speed (tail and centre) with reasonable accuracy.
For turbulent flows with free surfaces some important work involving LES with and without shear has been performed, as
follows.
Reboux et al. [321] used an LES approach for a turbulent interfacial two-phase flow. The effects of a modified Smagorinsky
model with and without shear interface treatment and the Variational Multiscale approach were examined. The numerical
results confirmed that both proposed methods were suitable for handling the anisotropy of turbulence in the liquid side
underneath the interfacial region. They also predict the formation of boundary layer stratification at the gas side. In addition,
turbulence statistics were calculated by both mentioned approaches, with better results for the Variational Multiscale
approach compared to the modified Smagorinsky model.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 725

Christensen [322] performed LES for the study of wave breaking and the interaction of the turbulence interface. A VOF
method was adopted for the simulation of the interface. Periodic spilling and plunging breakers were simulated by means
of LES along with the classical Smagorinsky model. The undertow, set-up and turbulence levels, obtained were compared
with experimental data. It was concluded that although the grid was relatively coarse, satisfactory results were obtained
for the wave height decay and undertow, while the turbulence levels were overestimated. The performance of the numerical
model was considered satisfactory.
The same year, Lubin et al. [323] conducted two-phase LES for the numerical simulation of plunging breaking waves in
air–water configurations, by using a dynamic (mixed scale) model by Sagaut [26] for the sub-grid scale modelling. In addi-
tion, they investigated the air entrainment mechanism that occurs during the wave breaking process. The numerical model
was validated with available experimental data and analytical solutions. In general, the proposed numerical model appeared
to be a reliable tool for the description of two-phase flows and gave very satisfactory results.
Lakehal and Liovic [324] undertook numerical simulations for the investigation of turbulence scales and their interaction
with wave breaking by means of Large-Eddy Interface Simulation (LEIS). LEIS is a combination of LES and an interface track-
ing method (e.g. VOF). The filtered single-fluid equations along with an ILES framework were employed to the MFVOF-3D
flow software. SGS modelling was treated by using the proposed SGS model of Liovic and Lakehal [325]. The numerical model
was validated with experimental (phase-averaged) data by Ting and Kirby [326] which are suitable for the validation of
breaking wave simulations. It was concluded that the numerical simulations were adequate in providing detailed informa-
tion for the air–liquid turbulent coherent structures, and their connexion with the local incidence transient mechanisms. For
the first time a conditional zonal analysis was undertaken for the investigation of transient mechanisms for turbulence sta-
tistics (turbulence kinetic energy, diffusion, decay and transport) and their effect and dependencies on the wave breaking
process.

4. Conclusions

This paper reviews the problem and successes of computing turbulent flows. The review is primarily concerned with the
most recent methods for such computer predictions. The successes and problems are demonstrated by listing and briefly
discussing several applications of DNS and LES to flows in pipes and free-surface flows.
In Table 4, a full list with the appropriate references of all the above mentioned turbulence strategies and models are
given, so that the potential user may easily find the appropriate information, for him to select the suitable turbulence model
for his own case of interest. The potential user is also recommended to examine the following review papers and guidelines
concerning verification and validation in CFD work, by Roache [327], Oberkampf and Trucano [328], Roy [329], Stern et al.
[330], ASME [331], NEA [332], ERCOFTAC [333], AIAA [334], in order to be able to assess the accuracy and reliability of his
numerical result.
The LES approach appears, from the given references that describe its applications, to have reached maturity. However,
further work is required to improve the characteristics of the method for more types of turbulent flow, in particular for com-
plex industrial flows. There are still challenges facing LES of turbulence such as the development of advanced sub-grid scale
models, high-order discretization techniques for eliminating the numerical errors, implementation on unstructured grids,
control of the numerical errors, interaction with other physical mechanisms and a ‘‘simple’’ wall stress model for wall-
bounded complex flows. In this context, the subject of ‘discretization techniques’ for convection is very important. Thus,
highly dissipative schemes must be avoided, as they smooth out several frequencies and therefore one may not attribute
errors correctly, either to the turbulence model or to the numerical scheme. On the other hand monotonic increasing
schemes appear beneficial (see MILES). This topic is also vast and, in several aspects, misunderstood, and is not covered
in the present work. The interested reader may find useful information in [335–339].
ILES techniques were developed recently and have gained a lot of application space due to the drawbacks of LES. Many
researchers replace the classical LES formulations, by introducing implicit SGS modelling with the use of nonlinear algo-
rithms (e.g. MILES). However, some challenges for MILES and in general for ILES, are the appropriate physical and mathemat-
ical framework for their analysis and development, and the interaction of the numerical schemes among the implicit SGS
models.
Hybrid LES/RANS methods are a good alternative to LES, because they combine the accuracy of LES and the speed of RANS.
These methods are suitable for flows dominated by large coherent structures and strong unsteady profiles (e.g. bluff body
flows, IC engines, among others). Hybrid LES/RANS methods along with the aforementioned PANS, PITM provide higher accu-
racy compared to the DES, SAS and URANS approaches, but they are also more demanding in CPU time.
The continuing progress of computer-hardware development is promising for DNS, in the near future, with major
improvements expected in statistical samples and in considering a variety of several physical parameters, for better under-
standing of the turbulence nature. Some final comments are:

– Some success has been achieved with two-equation models for relatively simple hydrodynamic phenomena; indeed, rou-
tine design work can now be undertaken in several applications of engineering practise, for which extensive studies have
optimised these models.
726 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

– Failures are still common for many applications, particularly those that involve strong curvature, intermittency, strong
buoyancy influences, low-Reynolds-number effects, rapid compression or expansion, strong swirl, and kinetically-influ-
enced chemical reaction. New conceptual developments are needed in these areas, probably along the lines of actually
calculating the principal manifestation of turbulence, e.g. intermittency. A start has been made in this direction in the
form of ‘multi-fluid’ models, and full simulations.
– Although some of the latest concepts hold promise of describing some of the most important physical consequences of
turbulence, they have not yet reached a definite stage of development.
– From this point of view, the older and simpler methods can still be recommended as the starting point (and sometimes
the finishing point) for engineering simulations.
– LES is currently the most accurate method available for practical computations and its use is expected to rise fast over the
next few years.
– DNS is obviously the method that provides the most precise and detailed description of turbulence but it is still out of
reach of the available everyday computer power, i.e. it cannot be used for everyday engineering design. It is, however,
even today very useful as it serves as a ‘‘test’’ of the other model predictions and of any new ideas on turbulence calcu-
lations. Until it becomes also a practical tool the authors recommend the use of ‘‘two-fluid’’ models that appear very
promising (but need further refinement), and LES along with its derivatives and the hybrid methods that have already
reached maturity.

Acknowledgements

The authors wish to thank Dr Djamel Lakehal, for kindly providing us with constructive remarks and recommendations on
an initial draft of this paper. Furthermore, special thanks are due to Dr. Stavros Karabelas and PhD Student Nektarios Koutso-
urakis for offering figures and useful material from their studies. The authors are indebted to the anonymous referees, for
extensively reviewing this paper and their useful comments. Finally, the author (C.A) wishes to express his sincere gratitude
to Prof. Omar Matar for the extension of his bursary and his encouragement during this study.
This work has been undertaken within the Joint Project on Transient Multiphase Flows and Flow Assurance. The Author
(C.A) wishes to acknowledge the contributions made to this project by the UK Engineering and Physical Sciences Research
Council (EPSRC) and the following: ASCOMP, GL Noble Denton, BP Exploration, CD-adapco, Chevron, ConocoPhillips, ENI,
ExxonMobil, FEESA, FMC Technologies, IFP Energies nouvelles, Granherne, Institutt for Energiteknikk, Kongsberg Oil & Gas
Technologies, MSi Kenny, PDVSA (INTEVEP), Petrobras, PETRONAS, SPT Group, Shell, SINTEF, Statoil and TOTAL.

References

[1] N.C. Markatos, The mathematical modelling of turbulent flows, Appl. Math. Modell. 10 (1986) 190–220.
[2] M. Hinze, T. Rung, Hydrodynamic shape-optimisation for turbulent industrial flows, Universitat Hamburg, Project C2, Poster presented at the
workshop Recent Trends and Future Developments in Computational Science and Engineering, Plon, 2014.
[3] H. Tennekes, J.L. Lumley, A First Course in Turbulence, The MIT Press, Cambridge, Massachusetts, 1972.
[4] J.C.R. Hunt, J.C. Vassilicos, Turbulence Structure and Vortex Dynamics, Cambridge University Press, 2000.
[5] C. Vassilicos, Intermittency in Turbulent Flows, Cambridge University Press, 2001.
[6] C.G. Speziale, Analytical methods for the development of Reynolds-stress closures in turbulence, Ann. Rev. Fluid Mech. 23 (1991) 107–157.
[7] H.K. Versteeg, W. Malalasekera, An Introduction to Computational Fluid Dynamics. The Finite Volume Method, second ed., Pearson Education Limited,
2007.
[8] W.C. Reynolds, T. Cebeci, Calculation of turbulent flows, in: P. Bradshaw (Ed.), Ch. 5 in Turbulence. Topics in Applied Physics, second ed., vol. 12,
Springer Verlag, Berlin, 1978, pp. 193–229.
[9] B.E. Launder, Heat and mass transport, in: P. Bradshaw (Ed.), Ch. 6 in Turbulence. Topics in Applied Physics, second ed., vol. 12, Springer Verlag, Berlin,
1978, pp. 231–287.
[10] P. Bradshaw, Introduction to turbulence, in: P. Bradshaw (Ed.), Ch. 1 in Turbulence. Topics in Applied Physics, second ed., vol. 12, Springer Verlag,
Berlin, 1978, pp. 1–44.
[11] K.R. Sreenivasan, R.A. Antonia, The phenomenology of small-scale turbulence, Annu. Rev. Fluid Mech. 29 (1997) 435–472.
[12] P.S. Klebanoff, NACA Report 1247, National Bureau of Standards, Washington D.C., 1995.
[13] F.M. White, Viscous Fluid Flow, second ed., McGraw Hill, New York, 1991.
[14] J. Paret, P. Tabeling, Intermittency in the two-dimensional inverse cascade of energy: experimental observations, Phys. Fluids 10 (1998) 3126–3136.
[15] M. Kholmyansky, A. Tsinober, On the origins of intermittency in real turbulent flows, in: C. Vassilicos (Ed.), Intermittency in Turbulent Flows,
Cambridge University Press, 2001, pp. 183–192.
[16] S. Tavoularis, U. Karnik, Further experiments on the evaluation of turbulent stresses and scales in uniformly sheared turbulence, J. Fluid Mech. 204
(1989) 457–478.
[17] R.S. Rogallo, P. Moin, Numerical simulation of turbulent flows, Annu. Rev. Fluid Mech. 16 (1984) 99–137.
[18] Z.-S. She, E. Jackson, S.A. Orszag, Structure and dynamics of homogeneous turbulence. models and simulations, Proc. R. Soc. London Ser. A Math. Phys.
Sci. 434 (1991) 101–124.
[19] L. Kleiser, T.A. Zang, Numerical simulation of transition in wall-bounded shear flows, Annu. Rev. Fluid Mech. 23 (1991) 495–537.
[20] N. Kasagi, N. Shikazono, Contribution of direct numerical simulation to understanding and modelling turbulent transport, Proc. R. Soc. London Ser. A
Math. Phys. Sci. 451 (1995) 257–292.
[21] P. Moin, K. Mahesh, Direct numerical simulation: a tool in turbulence research, Annu. Rev. Fluid Mech. 30 (1998) 539–578.
[22] R. Friedrich, T.J. Huttl, M. Manhart, C. Wagner, Direct numerical simulation of incompressible turbulent flows, Comput. Fluids 30 (2001) 555–579.
[23] T. Ishihara, T. Gotoh, Y. Kaneda, Study of high-Reynolds number isotropic turbulence by direct numerical simulation, Annu. Rev. Fluid Mech. 41 (2009)
165–180.
[24] S.B. Pope, Turbulent Flows, Cambridge University Press, 2000.
[25] L.C. Berselli, T. Iliescu, W.J. Layton, Mathematics of Large Eddy Simulation of Turbulent Flows, Springer, Berlin, New York, 2006.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 727

[26] P. Sagaut, Large Eddy Simulation for Incompressible Flows: An Introduction, third ed., Springer, Berlin (New York), 2006.
[27] F.F. Grinstein, L.G. Margolin, W.J. Rider, Implicit Large Eddy Simulation: Computing Turbulent Fluid Dynamics, Cambridge University Press,
Cambridge, UK, 2007.
[28] J. Meyers, B.J. Geurts, P. Sagaut, Quality and Reliability of Large-Eddy Simulations. ERCOFTAC Series, Springer, Berlin, New York, 2008.
[29] E. Garnier, N. Adams, P. Sagaut, Large Eddy Simulation for Compressible Flows, Springer, New York, 2009.
[30] X. Jiang, C.-H. Lai, Numerical Techniques for Direct and Large Eddy Simulations, Chapman & Hall/CRC, Boca Raton, 2009.
[31] M. Lesieur, O. Metais, New trends in large-eddy simulations of turbulence, Annu. Rev. Fluid Mech. 28 (1996) 45–82.
[32] U. Piomelli, Large-eddy simulation: achievements and challenges, Prog. Aeros. Sc. 35 (1999) 335–362.
[33] C. Meneveau, J. Katz, Scale-invariance and turbulence models for large-eddy simulation, Annu. Rev. Fluid Mech. 32 (2000) 1–32.
[34] S.B. Pope, Ten questions concerning the large-eddy simulation of turbulent flows, New J. Phys. 6 (2004) 35.
[35] C. Fureby, Towards the use of large eddy simulation in engineering, Prog. Aerosp. Sci. 44 (2008) 381–396.
[36] J. Frohlich, D. von Terzi, Hybrid LES/RANS methods for the simulation of turbulent flows, Prog. Aerosp. Sci. 44 (2008) 349–377.
[37] R. Bouffanais, Advances and challenges of applied large-eddy simulation, Comput. Fluids 39 (2010) 735–738.
[38] N.J. Georgiadis, D.P. Rizzetta, C. Fureby, Large-eddy simulation: current capabilities, recommended practices, and future research, AIAA J. 48 (2010)
1772–1784.
[39] P.G. Tucker, S. Lardeau, Applied large eddy simulation, Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 367 (2009) 2809–2818.
[40] T.B. Gatski, C.L. Rumsey, Linear and nonlinear eddy viscosity models, in: B.E. Launder, N.D. Sandham (Eds.), Closure Strategies for Turbulent and
Transitional Flows, Cambridge University Press, Cambridge, UK, 2002, pp. 9–46.
[41] ERCOFTAC, <www.ercoftac.org/special_interest_groups/15_turbulence_modelling/>.
[42] G.F. Hewitt, J.M. Delhaye, N. Zuber, Multiphase Science and Technology, vol. 5, CRC Press, 1989.
[43] M. Ishii, H. Takashi, Thermo-Fluid Dynamics of Two-Phase Flow, second ed., Springer-Verlag, Berlin, 2010.
[44] D.A. Drew, S.L. Passman, Theory of multicomponent FLUIDS, Applied Mathematical Sciences, vol. 135, Springer-Verlag, New York, 1998.
[45] P. Jenny, S.H. Lee, H.A. Tchelepi, Adaptive fully implicit multi-scale finite-volume method for multi-phase flow and transport in heterogeneous porous
media, J. Comput. Phys. 217 (2006) 627–641.
[46] D.B. Spalding, A general purpose computer program for multi-dimensional one- and two-phase flow, Math. Comput. Simul. 23 (1981) 267–276.
[47] S. Whitaker, The transport equations for multi-phase systems, Chem. Eng. Sci. 28 (1973) 139–147.
[48] B. Lecordier, D. Deware, L.M.J. Vervisch, Estimation of the accuracy of PIV treatments for turbulent flow studies by direct numerical simulations of
multi-phase flow, Meas. Sci. Technol. 12 (2001) 1382–1391.
[49] T. Berning, M. Odgaard, S.K. Kaer, A study of multi-phase flow through the cathode side of an interdigitated flow field using a multi-fluid model, J.
Power Sources 195 (2010) 4842–4852.
[50] D.A. Drew, Mathematical modeling of two-phase flow, Ann. Rev. Fluid Mech. 15 (1983) 261–291.
[51] A.E. Bergles, J.G. Collier, J.M. Delhaye, G.F. Hewitt, F. Mayinger, Two-phase Flow and Heat Transfer in the Power and Process Industries, Hemisphere
Publishing Corp., New York, 1981.
[52] D. Jamet, O. Lebaigue, N. Coutris, J.M. Delhaye, The second gradient method for the direct numerical simulation of liquid–vapor flows with phase
change, J. Comput. Phys. 169 (2001) 624–651.
[53] A. Prosperetti, Gretar Tryggvason, Computational Methods for Multiphase Flow, Cambridge University Press, 2007.
[54] N.C. Markatos, Modelling of two-phase transient flow and combustion of granular propellants, Int. J. Multiphase Flow 12 (1986) 913–933.
[55] D.K. Lilly, Numerical simulation of developing and decaying two-dimensional turbulence, J. Fluid Mech. 45 (1971) 395–415.
[56] S.A. Orszag, G.S. Patterson, Numerical simulation of three-dimensional homogeneous isotropic turbulence, Phys. Rev. Lett. 28 (1972) 76–79.
[57] R.S. Rogallo, Numerical experiments in homogeneous turbulence, Technical report, NASA TM 81315, 1981.
[58] J. Kim, P. Moin, R. Moser, Turbulence statistics in fully developed channel flow at low Reynolds number, J. Fluid Mech. 177 (1987) 133–166.
[59] R.D. Moser, J. Kim, N.N. Mansour, Direct numerical simulation of turbulent channel flow up to Res = 590, Phys. Fluids 11 (1999) 943–945.
[60] H. Abe, H. Kawamura, Y. Matsuo, Direct numerical simulation of a fully developed turbulent channel flow with respect to the Reynolds umber
dependence, J. Fluids Eng. 123 (2001) 382–393.
[61] J.C. Del Alamo, J. Jimenez, P. Zandonade, R.D. Moser, Scaling of the energy spectra of turbulent channels, J. Fluid Mech. 500 (2004) 135–144.
[62] S. Hoyas, J. Jimenez, Scaling of the velocity fluctuations in turbulent channels up to Res = 2003, Phys. Fluids 18 (2006) 011702.
[63] J. Komminaho, M. Skote, Reynolds stress budgets in couette and boundary layer flows, Flow Turbul. Combust. 68 (2002) 167–192.
[64] P.R. Spalart, Direct simulation of a turbulent boundary layer up to Reh = 1410, J. Fluid Mech. 187 (1988) 61–98.
[65] P. Schlatter, R. Orlu, Q. Li, G. Brethouwer, J.H.M. Fransson, A.V. Johansson, P.H. Alfredsson, D.S. Henningson, Turbulent boundary layers up to
Reh = 2500 studied through simulation and experiment, Phys. Fluids 21 (2009). 051702–4.
[66] A. Ferrante, S. Elghobashi, Reynolds number effect on drag reduction in a microbubble-laden spatially developing turbulent boundary layer, J. Fluid
Mech. 543 (2005) 93–106.
[67] P. Schlatter, R. Orlu, Assessment of direct numerical simulation data of turbulent boundary layers, J. Fluid Mech. 659 (2010) 116–126.
[68] J. Sillero, J. Jimenez, R.D. Moser, N.P. Malaya, Direct simulation of a zero-pressure-gradient turbulent boundary layer up to Reh = 6650, J. Phys: Conf.
Ser. 318 (2011) 022023.
[69] J.A. Sillero, J. Jimenez, R. Moser, One-point statistics for turbulent wall-bounded flows at Reynolds number up to + 2000, Phys. Fluids 25 (2013)
105102.
[70] G. Alfonsi, On direct numerical simulation of turbulent flows, Appl. Mech. Rev 64 (2011) 1–33.
[71] M. Lee, N. Malaya, R.D. Moser, Petascale direct numerical simulation of turbulent channel flow on up to 786K cores, in: Proceedings of the
International Conference on High Performance Computing, Networking, Storage and Analysis, Denver, Colorado, 2013.
[72] Y. Kaneda, T. Ishihara, M. Yokokawa, K. Itakura, A. Uno, Energy dissipation rate and energy spectrum in high resolution direct numerical simulations of
turbulence in a periodic box, Phys. Fluids 15 (2003) L21–L24.
[73] Y. Kaneda, T. Ishihara, High-resolution direct numerical simulation of turbulence, J. Turbul. 7 (2006) 1–17.
[74] W.C. Reynolds, The potential and limitations of direct and large eddy simulations, in: John L. Lumley (Ed.), Whither Turbulence? Turbulence at the
Crossroads, Lecture Notes in Physics, 357, Springer-Verlag, NY, 1989, pp. 313–342.
[75] N.D. Sandham, Introduction to direct numerical simulation, in: B.E. Launder, N.D. Sandham (Eds.), Closure Strategies for Turbulent and Transitional
Flows, Cambridge University Press, Cambridge, UK, 2002.
[76] J. Jimenez, P. Moin, The minimal flow unit in near-wall turbulence, J. Fluid Mech. 225 (1991) 213–240.
[77] H. Choi, P. Moin, J. Kim, Active turbulence control for drag reduction in wall-bounded flows, J. Fluid Mech. 262 (1994) 75–110.
[78] G.N. Coleman, R.D. Sandberg, A primer on direct numerical simulation of turbulence-methods, procedures and guidelines. Tech. Rep AFM-09/01a.
Aerodynamics & Flight Mechanics Research Group, School of Engineering Sciences University of Southampton, Southampton, UK, 2010.
[79] N.C. Markatos, Computer simulation techniques for turbulent flows, Encycl. Fluid Mech. 6 (1) (1985) 1259–1275.
[80] N.C. Markatos, Dynamic computer modelling of environmental systems for decision making, risk assessment and design, Asia-Pac. J. Chem. Eng. 7
(2012) 182–205.
[81] P. Bradshaw, An Introduction to Turbulence and its Measurements, Pergamon, Oxford, 1971.
[82] W. Frost, J. Bitte, Statistical Concepts in Turbulence, Handbook of Turbulence, vol. 1, Plenum Press, New York, 1977.
[83] L.S.G. Kovasznay, Turbulent shear flows, Convegno Sulla Teoria Della Turbulenza, Rome, 1970.
[84] K. Hanjalic, One-point closure models for buoyancy-driven turbulent flows, Annu. Rev. Fluid Mech. 34 (2002) 321–347.
[85] K. Hanjalic, Will RANS survive LES? a view of perspectives, J. Fluids Eng. 127 (2005) 831–839.
728 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

[86] D. Wilcox, Turbulence Modelling for CFD, third ed., DCW Industries, Inc, 2006.
[87] W. Rodi, Turbulence models and their application in hydraulics, A state-of-the-art review, in: IAHR Monographs, third ed., CRC Press, 1993.
[88] L. Prandtl, Report on investigation of developed turbulence, ZAMM 5 (1925) 136–139.
[89] T. Cebeci, A.M.O. Smith, Analysis of Turbulent Boundary Layers, Academic Press, New York, 1974.
[90] B.S.Baldwin, H. Lomax, Thin layer approximation and algebraic model for separated turbulent flows, AIAA 16 Aerospace Sciences Meeting, 1978.
[91] K. Hanjalic, Closure models for incompressible turbulent flows (lecture series), in: J.P.A.J. Beeck, C. Benocci (Eds.), Introduction to Turbulence
Modelling, Von Kaman Institute for Fluid Dynamics, Belgium, 2004, pp. 1–75.
[92] D.A. Johnson, L.S. King, A mathematically simple turbulence closure model for attached and separated turbulent boundary layers, AIAA J. 23 (1985)
1684–1692.
[93] D.A. Johnson, Transonic separated flow predictions with an eddy-viscosity/-Reynolds-stress closure model, AIAA J. 25 (1987) 252–259.
[94] D.A. Johnson, T.J. Coakley, Improvements to a non-equilibrium algebraic turbulence model, AIAA J. 28 (1990) 2000–2003.
[95] A.M. Savill, T.B. Gatski, P.A. Lindberg, A pseudo-3D extension to the Johnson-King model and its application to the EuropeExpt S-duct, in: O.
Pironneau, W. Rodi, I.L. Rhyming, A.M. Savill, T.V. Truong (Eds.) Numerical Simulation of Unsteady Flows and Transitions to Turbulence, Cambridge,
1992, pp 158–165.
[96] B.S. Baldwin, T.J. Barth, A one-equation turbulence transport model for high-Reynolds number wall-bounded flows. AIAA, Aerospace Sciences
Meeting, 1991.
[97] P.R. Spalart, S.R. Allmaras, A one-equation turbulence model for aerodynamics flows, La Recherche Aerospatiale 1 (1994) 5–21.
[98] P.R. Spalart, M. Shur, On the sensitization of turbulence models to rotation and curvature, Aerosp. Sci. Technol. 5 (1997) 297–302.
[99] J. Dacles-Mariani, G.G. Zilliac, J.S. Chow, P. Bradshaw, Numerical/experimental study of a wingtip vortex in the near field, AIAA J. 33 (1995) 1561–
1568.
[100] M.M. Rahman, T. Siikonen, R.K. Agarwal, Improved low-Reynolds-number one-equation turbulence model, AIAA J. 49 (2011) 735–747.
[101] S.J. Karabelas, N.C. Markatos, Water vapor condensation in forced convection flow over an airfoil, Aerosp. Sci. Tech. 12 (2008) 150–158.
[102] S.J. Karabelas, High Reynolds number flow past a flapping multi-element airfoil, in: 43 AIAA Fluid Dynamics Conference, San Diego, USA, 2013.
[103] S.M. Klausmeyer, J.C. Lin, Comparative results from a CFD challenge over a 2D three-element high-lift airfoilm, NASA TM, 1997.
[104] J.C. Lin, C.J. Dominik, Parametric investigation of a high-lift airfoil at high Reynolds numbers, J. Aircr. 34 (1997) 485–491.
[105] E. Fares, W. Schroder, A general one-equation turbulence model for free shear and wall-bounded flows, Flow Turbul. Combust. 73 (2004) 187–215.
[106] G. Alfonsi, Reynolds-averaged Navier–Stokes equations for turbulence modelling, Appl. Mech. Rev. 62 (2009) 1–20.
[107] B.E. Launder, B.I. Sharma, Application of the energy dissipation model of turbulence to the calculation of flow near a spinning disk, Lett. Heat Mass
Transfer 1 (1974) 131–138.
[108] D.C. Wilcox, Reassessment of the scale-determining equation for advanced turbulence models, AIAA J. 26 (1988) 1299–1310.
[109] T.-H. Shih, W.W. Liou, A. Shabbir, Z. Yang, J. Zhu, A new k–e eddy viscosity model for high Reynolds number turbulent flows, Comput. Fluids 24 (1995)
(1995) 227–238.
[110] V. Yakhot, S.A. Orszag, S. Thangam, T.B. Gatski, C.G. Speziale, Development of turbulence models for shear flows by a double expansion technique,
Phys. Fluids A Fluid Dyn. 4 (1992) 1510–1520.
[111] C.D. Argyropoulos, M.N. Christolis, Z. Nivolianitou, N.C. Markatos, Assessment of acute effects for fire-fighters during a fuel-tank fire, in: International
Conference Working On Safety Net, Prevention of Occupational Accident in a Changing Work Environment, Crete, Greece, 2008.
[112] N.C. Markatos, C. Christolis, C. Argyropoulos, Mathematical modelling of toxic pollutants dispersion from large tank fires and assessment of acute
effects for fire fighters, Int. J. Heat Mass Transfer 52 (2009) 4021–4030.
[113] C.D. Argyropoulos, G.M. Sideris, M.N. Christolis, Z. Nivolianitou, N.C. Markatos, Modelling pollutants dispersion and plume rise from large
hydrocarbon tank fires in neutrally stratified atmosphere, Atmos. Environ. 44 (2010) 803–813.
[114] C.D. Argyropoulos, M.N. Christolis, Z. Nivolianitou, N.C. Markatos, A hazards assessment methodology for large liquid hydrocarbon fuel tanks, J. Loss
Prev. Process. Ind. 25 (2012) 329–335.
[115] N. Koutsourakis, J. Bartzis, N. Markatos, Evaluation of Reynolds stress, k–e and RNG k–e turbulence models in street canyon flows using various
experimental datasets, Environ. Fluid Mech. 12 (2012) 379–403.
[116] S.J. Karabelas, B.C. Koumroglou, C.D. Argyropoulos, N.C. Markatos, High Reynolds number turbulent flow past a rotating cylinder, Appl. Math. Modell.
36 (2012) 379–398.
[117] A.N. Kolmogorov, Equations of turbulent motion of an incompressible fluid, Izv Acad. Sci. USSR Phys. 6 (1942) 56–58.
[118] B.E. Launder, D.B. Spalding, Mathematical Models of Turbulence, Academic Press, London, 1972.
[119] F.R. Menter, Review of the shear-stress transport turbulence model experience from an industrial perspective, Int. J. Comput. Fluid Dyn. 23 (2009)
305–316.
[120] F.R. Menter, Two-equation eddy-viscosity turbulence models for engineering applications, AIAA J. 32 (1994) 1598–1605.
[121] P. Smirnov, F.R. Menter, Sensitizing of the SST turbulence model to rotation and curvature by applying the Spalart-Shur correction term, in:
Proceedings of ASME Turbo Expo, Berlin, 2008.
[122] K. Hanjalic, B.E. Launder, R. Schiestel, Multiple-time-scale concepts in turbulent transport modelling, in: L.J.S. Bradbury et al. (Eds.), Turbulent Shear
Flows, 2, Springer-Verlag, Berlin, 1980, pp. 36–49.
[123] O. Heynes, M. Cotton, T. Craft, Eddy-viscosity and stress-transport turbulence models in application to a plane synthetic jet, Flow Turbul. Combust. 91
(2013) 931–947.
[124] Q. Zhao, S. Armfield, K. Tanimoto, Numerical simulation of breaking waves by a multi-scale turbulence model, Coast. Eng. 51 (2004) 53–80.
[125] O. Gregoire, D. Souffland, S. Gauthier, R. Schiestel, A two-time-scale turbulence model for compressible flows: turbulence dominated by mean
deformation interaction, Phys. Fluids 11 (1999) 3793–3807.
[126] K.R. Kim, M.A. Cotton, T.J. Craft, O.R. Heynes, On the dynamics and frequency response of fully-pulsed turbulent round jets: computations using two-
time-scale/strain sensitized eddy viscosity models, Int. J. Heat Fluid Flow 29 (2008) 1650–1669.
[127] T.S. Klein, T.J. Craft, H. Iacovides, Development in two-time-scale turbulence models applied to non-equilibrium flows, TSFP-7, 2011.
[128] V.C. Patel, W. Rodi, G. Scheuerer, Turbulence models for near-wall and low Reynolds number flows – a review, AIAA J. 23 (1985) (1985) 1308–1319.
[129] C.K.G. Lam, K.A. Bremhorst, Modified form of k–e model for predicting wall turbulence, ASME J. Fluids Eng. 103 (1981) 456–460.
[130] J. Bredberg, S.-H. Peng, L. Davidson, An improved k–x turbulence model applied to recirculating flows, Int. J. Heat Fluid Flow 23 (2002) 731–743.
[131] D.C. Wilcox, Simulation of transition with a two-equation turbulence model, AIAA J. 32 (1994) 247–255.
[132] S.H. Peng, L. Davidson, S. Holmberg, A modified low-Reynolds-number k–x model for recirculating flows, J. Fluid Eng. 119 (1997) 867–875.
[133] P.R. Resende, F.T. Pinho, B.A. Younis, K. Kim, R. Sureshkumar, Development of a low-Reynolds-number k–x model for FENE-P fluids, Flow Turbul.
Combust. 90 (2013) 69–94.
[134] T.B. Gatski, C.G. Speziale, On explicit algebraic stress models for complex turbulent flows, J. Fluid Mech. 254 (1993) 59–78.
[135] T. Shih, J. Zhu, J. Lumley, A new Reynolds stress algebraic equation model, Comput. Meth. Appl. Mech. Eng. 125 (1995) 287–302.
[136] A. Hellsten, S. Wallin, Explicit algebraic Reynolds stress and non-linear eddy-viscosity models, Int. J. Comput. Fluid Dyn. 23 (2009) 349–361.
[137] T.J. Craft, B.E. Launder, K. Suga, Prediction of turbulent transitional phenomena with a nonlinear eddy-viscosity model, Int. J. Heat Fluid Flow 18
(1997) 15–28.
[138] D.D. Apsley, M.A. Leschziner, A new low-Reynolds-number nonlinear two-equation turbulence model for complex flows, Int. J. Heat Fluid Flow 19
(1988) 209–222.
[139] P.A. Durbin, Near-wall turbulence closure modelling without ‘‘damping functions’’, Theor. Comput. Fluid Dyn. 3 (1991) 1–13.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 729

[140] K. Hanjalic, M. Popovac, M. Hadziabdic, A robust near-wall elliptic relaxation eddy-viscosity turbulence model for CFD, Int. J. Heat Fluid Flow 25
(2004) 1047–1051.
[141] P.A. Durbin, B.A. Pettersson, The elliptic relaxation method, in: B.E. Launder, N.D. Sandham (Eds.), Closure Strategies for Turbulent and Transitional
Flows, Cambridge University Press, Cambridge, UK, 2002, pp. 127–152.
[142] P.A. Durbin, On the k–e stagnation point anomaly, Int. J. Heat Fluid Flow 17 (1996) 89–90.
[143] F. Billard, D. Laurence, A robust k–e–t2/k elliptic blending turbulence model applied to near-wall, separated and buoyant flows, Int. J. Heat Fluid Flow
33 (2012) 45–58.
[144] R. Manceau, K. Hanjalic, Elliptic blending model: a new near-wall Reynolds-stress turbulence closure, Phys. Fluids 14 (2002) 744–754.
[145] F. Billard, A. Revell, T. Craft, Application of recently developed elliptic blending based models to separated flows, Int. J. Heat Fluid Flow 35 (2012) 141–
151.
[146] B.J. Daly, F.H. Harlow, Transport equations in turbulence, Phys. Fluids 13 (1970) 2634–2649.
[147] J. Magnaudet, The modelling of inhomogeneous turbulence in the absence of mean velocity gradients, Fourth Eur. Turbulence Conf., Delft, The
Netherlands, 1992.
[148] Y. Nagano, M. Tagawa, Turbulence models for triple velocity and scalar correlations, Turbulent Shear Flows, vol. 7, Springer, Berlin, 1991, pp. 63–78.
[149] K. Hanjalic, Advanced turbulence closure models: a view of current status and future prospects, Int. J. Heat Fluid Flow 15 (1994) 178–203.
[150] D. Naot, A. Shavit, M. Wolfshtein, Two-point correlation model and the redistribution of Reynolds stresses, Phys. Fluids 16 (1973) 738–743.
[151] B.E. Launder, G.J. Reece, W. Rodi, Progress in the development of a Reynolds-stress turbulence closure, J. Fluid Mech. 68 (1975) 537–566.
[152] C.G. Speziale, S. Sarkar, T.B. Gatski, Modelling the pressure-strain correlation of turbulence: an invariant system dynamic approach, J. Fluid Mech. 227
(1991) 245–272.
[153] T.J. Craft, B.E. Launder, A Reynolds stress closure designed for complex geometries, Int. J. Heat Fluid Flow 17 (1996) 245–254.
[154] B.E. Launder, D.P. Tselepidakis, Application of a new second-moment closure to turbulent channel flow rotating in orthogonal mode, Int. J. Heat Fluid
Flow 15 (1994) 2–10.
[155] K. Hanjalic, S. Jakirlic, Second-moment turbulence closure modelling, in: B.E. Launder, N.D. Sandham (Eds.), Closure Strategies for Turbulent and
Transitional Flows, Cambridge University Press, Cambridge, UK, 2002, pp. 47–101.
[156] A. Kovar-Panskus, P. Louka, J.F. Sini, E. Savory, M. Czech, A. Abdelqari, P.G. Mestayer, N. Toy, Influence of geometry on the mean flow within urban
street canyons—a comparison of wind tunnel experiments and numerical simulations, Water Air Soil Pollut. 2 (2002) 365–380.
[157] X.X. Li, D.Y.C. Leung, C.H. Liu, K.M. Lam, Physical modeling of flow field inside urban street canyons, J. Appl. Meteorol. Clim. 47 (2008) 2058–2067.
[158] X.X. Li, C.H. Liu, D.Y.C. Leung, Large-eddy simulation of flow and pollutant dispersion in high-aspect-ratio urban street canyons with wall model,
Bound-Layer Meteorol. 129 (2008) 249–268.
[159] P.A. Durbin, Reynolds stress model for near-wall turbulence, J. Fluid Mech. 249 (1993) 465–498.
[160] K. Hanjalic, S. Kenjeres, Some developments in turbulence modelling for wind and environmental engineering, J. Wind Eng. Ind. Aerodyn. 96 (2008)
1537–1570.
[161] L. Thielen, H.J.J. Jonker, K. Hanjalic, Symmetry breaking of flow and heat transfer in multiple impinging jets, Int. J. Heat Fluid Flow 24 (2003) 444–453.
[162] B. Basara, S. Jakirlic, A new hybrid turbulence modelling strategy for industrial CFD, Int. J. Numer. Methods Fluids 42 (2003) 89–116.
[163] S.B. Pope, A more general effective-viscosity hypothesis, J. Fluid Mech. 72 (1975) 331–340.
[164] T. Jongen, T.B. Gatski, General explicit algebraic stress relations and best approximation for three-dimensional flows, Int. J. Eng. Sci. 36 (1998) 739–
763.
[165] S. Wallin, A.V. Johansson, An explicit algebraic Reynolds stress model for incompressible and compressible turbulent flows, J Fluid Mech. 403 (2000)
89–132.
[166] D.B. Spalding, Chemical reaction in turbulent fluids, J. Physico-chemical Hydrodyn. 4 (1983) 323–336.
[167] M.R. Malin, D.B. Spalding, A two-fluid model of turbulence and its application to heated plane jets and wakes, J. Physico-chemical Hydrodyn. 5 (1984)
339–362.
[168] N.C. Markatos, C.A. Kotsifaki, One-dimensional, two-fluid modelling of turbulent premixed flames, Appl. Math. Modell. 18 (1994) 646–657.
[169] Y.M. Shen, C.-O. Ng, A.T. Chwang, A two-fluid model of turbulent two-phase flow for simulating turbulent stratified flows, Ocean Eng. 30 (2003) 153–
161.
[170] K.Z. Yu, G.-L. Ding, T.-J. Chen, Modified two-fluid model for air curtains in open vertical display cabinets, Int. J. Refrig. 31 (2008) 472–482.
[171] D.L. Youngs, Representation of the molecular mixing process in a two-phase flow turbulent mixing model, in: Proc. Fifth Int. Workshop on the Physics
of Compressible Turbulent Mixing, World Scientific, Singapore, 1995, pp. 83–88.
[172] A. Llor, P. Bailly, A new turbulent two-field concept for modelling Rayleigh–Taylor, Richtmyer–Meshkov, and Kelvin–Helmholtz mixing layers, Lasers
Beams 21 (2003) 311–315.
[173] D. Liu, C. Wu, K. Linden, J. Ducoste, Numerical simulation of UV disinfection reactors: evaluation of alternative turbulence models, Appl. Math. Modell.
31 (2007) 1753–1769.
[174] Z. Cao, H. Han, B. Gu, A novel optimization strategy for the design of air curtains for open vertical refrigerated display cases, Appl. Therm. Eng. 31
(2011) 3098–3105.
[175] H. Tinoco, H. Lindqvist, W. Frid, Numerical simulation of industrial flows, numerical simulations – examples and applications in computational
fluid dynamics, Prof. Lutz Angermann (Ed.), ISBN: 978-953-307-153-4, InTech, DOI: http://dx.doi.org/10.5772/13216, 2010. Available from:
<http://www.intechopen.com/books/numerical-simulations-examples-and-applications-in-computational-fluid-dynamics/numerical-simulation-of-
industrial-flows>.
[176] M. Meinke, Th. Rister, F. Rutten, A. Schvorack, Simulation of internal and free turbulent flows, in: H.J. Bungartz, F. Durst, C. Zenger (Eds.), High
Performance Scientific and Engineering Computing, Springer-Verlag, 1998, pp. 61–79.
[177] J. Frohlich, W. Rodi, Introduction to large eddy simulation of turbulent flows, in: B.E. Launder, N.D. Sandham (Eds.), Closure Strategies for Turbulent
and Transitional Flows, Cambridge University Press, Cambridge, UK, 2002, pp. 267–298.
[178] J. Smagorinsky, General circulation experiments with the primitive equations, Mon. Weather Rev. 91 (1963) 99–164.
[179] D.K. Lilly, The representation of small-scale turbulence in numerical simulation experiments, in: Proceedings IBM Scientific Computing Symposium
on Environmental Sciences, 1967, pp. 195–210.
[180] J.W. Deardorff, A numerical study of three-dimensional turbulent channel flow at large Reynolds numbers, J. Fluid Mech. 41 (1970) 453–480.
[181] R.H. Kraichnan, Eddy viscosity in two and three dimensions, J. Atmos. Sci. 33 (1976) 1521–1536.
[182] J.R. Chasnov, Simulation of the Kolmogorov inertial subrange using an improved subgrid model, Phys. Fluids A 3 (1991) 188–200.
[183] U. Schumann, Subgrid scale model for finite difference simulations of turbulent flows in plane channels and annuli, J. Comput. Phys. 18 (1975) 376–
404.
[184] P. Moin, J. Kim, Numerical investigation of turbulent channel flow, J. Fluid Mech. 118 (1982) 341–377.
[185] U. Piomelli, High Reynolds number calculations using the dynamic subgrid-scale stress model, Phys. Fluids A 5 (1993) 1484–1490.
[186] A. Leonard, Energy cascade in large-eddy simulations of turbulent fluid flows, Proceedings of Turbulent Diffusion in Environmental Pollution,
Charlottesville, vol. 18, Elsevier, 1974, pp. 237–248.
[187] A. Aldama, Filtering techniques for turbulent flow simulations, Lecture Notes in Engineering, vol. 56, Springer, Berlin, 1990.
[188] R.A. Clark, J.H. Ferziger, W.C. Reynolds, Evaluation of subgrid-scale models using an accurately simulated turbulent flow, J. Fluid Mech. 91 (1979) 1–
16.
[189] J.P. Boris, F.F. Grinstein, E.S. Oran, R.L. Kolbe, New insights into large eddy simulation, Fluid Dyn. Res. 10 (1992) 199–228.
[190] S.J. Karabelas, Large eddy simulation of high-Reynolds number flow past a rotating cylinder, Int. J. Heat Fluid Flow 31 (2010) 518–527.
730 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

[191] M. Germano, U. Piomelli, P. Moin, W.H. Cabot, A dynamic subgrid scale eddy viscosity model, Phys. Fluids A 3 (1991) 1760–1765.
[192] D.K. Lilly, A proposed modification of the Germano subgrid-scale closure method, Phys. Fluids A 4 (1992) 633–635.
[193] S. Ghosal, T.S. Lund, P. Moin, K. Akselvoll, A dynamic localization model for large-eddy simulation of turbulent flows, J. Fluid Mech. 286 (1995) 229–
255.
[194] C. Meneveau, T.S. Lund, W.H. Cabot, A lagrangian dynamic subgrid-scale model of turbulence, J. Fluid Mech. 319 (1996) 353–385.
[195] J. Bardina, J.H. Ferziger, W.C. Reynolds, Improved subgrid scale models for large eddy simulation, in: AIAA 13th Fluid and Plasma Dynamics
Conference, Snowmass, Colorado, 1980.
[196] A. Leonard, Large-eddy simulation of chaotic convection and beyond, Tech. Rep. 97–0204 AIAA, 1997.
[197] B. Geurts, Inverse modelling for large eddy simulation, Phys. Fluids 9 (1997) 3585–3587.
[198] Y. Zang, R.L. Street, J.R. Koseff, A dynamic mixed subgrid scale model and its application to turbulent recirculating flows, Phys. Fluid 5 (1993) 3186–
3196.
[199] B. Vreman, B. Geurts, H. Kuerten, On the formulation of the dynamic mixed subgrid-scale model, Phys. Fluid 6 (1994) 4057–4059.
[200] A. Leonard, G.S. Winckelmans, A tensor-diffusivity subgrid model for large-eddy simulation, in: P.R. Voke, N.D. Sandham, L. Kleiser (Eds.), Direct and
Large-Eddy Simulation III, 1999, pp. 147–162.
[201] S. Stolz, N.A. Adams, An approximate deconvolution procedure for large-eddy simulation, Phys. Fluids 11 (1999) 1699–1701.
[202] S. Stolz, N.A. Adams, L. Kleiser, The approximate deconvolution model for LES of compressible flows and its application to shock-turbulent-boundary-
layer interaction, Phys. Fluids 13 (2001) 2985–3001.
[203] S. Stolz, N.A. Adams, L. Kleiser, An approximate deconvolution model for large eddy simulations with application to incompressible wall-bounded
flows, Phys. Fluids 13 (2001) 997–1015.
[204] N.A. Adams, S. Hickel, T. Kempe, J.A. Domaradzki, On the relation between subgrid-scale modelling and numerical discretization in large-eddy
simulation, in: Complex Effects in Large Eddy Simulations, Springer, Berlin Heidelberg, 2007, pp. 15–27.
[205] B.J. Geurts, Elements of direct and large-eddy simulation, R.T. Edwards Inc, Flourtown, 2003.
[206] O. Metais, M. Lesieur, Spectral large-eddy simulation of isotropic and stably stratified turbulence, J. Fluid Mech. 239 (1992) 157–194.
[207] F. Ducros, P. Comte, M. Lesieur, Large-eddy simulation of transition to turbulence in a boundary layer developing spatially over a flat plate’’, J. Fluid
Mech. 326 (1996) 1–36.
[208] J.A. Domaradzki, E.M. Saika, A subgrid-scale model based on the estimation of unresolved scales of turbulence, Phys. Fluids 9 (1997) 2148–2164.
[209] J.A. Domaradzki, K.C. Loh, The subgrid-scale estimation model in the physical space representation, Phys. Fluids 11 (1999) 2330–2342.
[210] C. Cao, D. Holm, E.S. Titi, On the clark-a model of turbulence: global regularity and long-time dynamics, J. Turbul. 6 (2005) 1–11.
[211] A. Cheskidov, D.D. Holm, E. Olson, E.S. Titi, On a Leray-a model of turbulence, Proc. R. Soc. A Math. Phys. Eng. Sci. 461 (2005) 629–649.
[212] S. Chen, C. Foias, D.D. Holm, E. Olson, E.S. Titi, S. Wynne, The Camassa–Holm equations and turbulence, Phys. D 133 (1999) 49–65.
[213] S. Chen, C. Foias, D.D. Holm, E. Olson, E.S. Titi, S. Wynne, A connection between the Camassa–Holm equations and turbulent flows in channels and
pipes, Phys. Fluids 11 (1999) 2343–2353.
[214] S. Chen, C. Foias, D.D. Holm, E. Olson, E.S. Titi, S. Wynne, Camassa–Holm equations as a closure model for turbulent channel and pipe flow, Phys. Rev.
Lett. 81 (1998) 5338–5341.
[215] B.J. Geurts, D.D. Holm, Regularization modelling for large-eddy simulation, Phys. Fluids 15 (2003) 13–16.
[216] J. Leray, Essai sur le mouvement d’un fluide visqueux emplissant l’space, Acta Math. 63 (1934) 193–248.
[217] A.A. Llyin, E.M. Lunasin, E.S. Titi, A modified-Leray-a subgrid scale model of turbulence, Nonlinearity 19 (2006) 879–897.
[218] U. Piomelli, Large-eddy and direct simulation of turbulent flows, in: CFD2001 – 9e conférence annuelle de la société Canadienne de CFD, Kitchener,
Ontario, Canada, 2001.
[219] U. Piomelli, E. Balaras, Wall-layer models for large-eddy simulations, Annu. Rev. Fluid Mech. 34 (2002) 349–374.
[220] B.J. Geurts, A. Leonard, Is LES ready for complex flows?, in: B.E. Launder, N.D. Sandham (Eds.), Closure Strategies for Turbulent and Transitional Flows,
Cambridge University Press, Cambridge, UK, 2002, pp. 720–739.
[221] B.J. Geurts, Interacting errors in large-eddy simulation: a review of recent developments, J. Turbul. 7 (2006) 1–16.
[222] B.J. Geurts, Analysis of errors occurring in large eddy simulation, Proc. R. Soc. A Math. Phys. Eng. Sci. 367 (2009) 2873–2883.
[223] I. Celik, M. Klein, M. Freitag, J. Janicka, Assessment measures for URANS/DES/LES: an overview with applications, J. Turbul. 7 (2006) 1–27.
[224] B.J. Geurts, J. Frohlich, A framework for predicting accuracy limitations in large-eddy simulations, Phys. Fluids 14 (2002) 41–44.
[225] J. Meyers, B.J. Geurts, M. Baelmans, Database analysis of errors in large eddy simulation, Phys. Fluids 15 (2003) 2740–2755.
[226] M. Freitag, M. Klein, An improved method to assess the quality of large eddy simulations in the context of implicit filtering, J. Turbul. 7 (2006) 1–11.
[227] D. Drikakis, W. Rider, High Resolution Methods for Incompressible and Low-speed Flows, Springer-Verlag, Berlin Heidelberg, 2005.
[228] J.P. Boris, D.L. Book, Flux-corrected transport. I. SHASTA, a fluid transport algorithm that works, J. Comput. Phys. 11 (1973) 38–69.
[229] B. van Leer, Towards the ultimate conservative difference scheme. I. The quest of monotonicity, in: H. Cabannes, R. Temam (Eds.), Lecture Notes in
Physics, vol. 18, Springer-Verlag, Berlin, 1973, pp. 163–168.
[230] P. Colella, P.R. Woodward, The piecewise parabolic method (PPM) for gas-dynamical simulations, J. Comput. Phys. 54 (1984) 174–201.
[231] A. Harten, High resolution schemes for hyperbolic conservation laws, J. Comput. Phys. 49 (1983) 357–393.
[232] P. Colella, A direct Eulerian MUSCL scheme for gas dynamics, SIAM. J. Sci. Statist. Comput. 6 (1985) 104–117.
[233] C. Fureby, M. Liefvendahl, U. Svennberg, L. Persson, T. Persson, Incompressible wall-bounded flows, in: F.F. Grinstein, L. Margolin, B. Rider (Eds.),
Implicit Large Eddy Simulation: Computing Turbulent Fluid Dynamics, Cambridge University Press, 2007, pp. 301–328.
[234] G. Patnaik, F.F. Grinstein, J.P. Boris, T.R. Young, O. Parmhed, Large-scale urban simulations, in: F.F. Grinstein, L. Margolin, B. Rider (Eds.), Implicit Large
Eddy Simulation: Computing Turbulent Fluid Dynamics, Cambridge University Press, 2007, pp. 502–530.
[235] J.P. Boris, On large eddy simulation using subgrid turbulence models, comment 1, in wither turbulence? Turbulence at the crossroads, Lect. Notes
Phys. 357 (1990) 344–353.
[236] J.P. Boris, More for LES: a brief historical perspective of MILES, in: F.F. Grinstein, L. Margolin, B. Rider (Eds.), Implicit Large Eddy Simulation:
Computing Turbulent Fluid Dynamics, Cambridge University Press, 2007, pp. 9–38.
[237] S. Johansson, L. Davidson, E. Olsson, Numerical simulation of vortex shedding past triangular cylinders at high Reynolds numbers using a k–e
turbulence model, Int. J. Numer. Methods Fluids 16 (1993) 859–878.
[238] S. Perzon, L. Davidson, On CFD and transient flow in vehicle aerodynamics, in: SAE Technical Paper 2000-01-0873, Detroit, 2000.
[239] S. Perzon, L. Davidson, On transient modelling of the flow around vehicles using the Reynolds equation, in: J.-H. Wu, Z.-J. Zhu, F.-P. Jia, X.-B. Wen, W.
Hu (Eds.), ACFD 2000, Bejing, 2000.
[240] C. Speziale, Turbulence modelling for time-dependent RANS and VLES: a review, AIAA J. 36 (1998) 173–184.
[241] P.R. Spalart, Strategies for turbulence modelling and simulations, Int. J. Heat Fluid Flow 21 (2000) 252–263.
[242] S. Kenjeres, K. Hanjalic, Tackling complex turbulent flows with transient RANS, Fluid Dyn. Res. 41 (2009) 1–32.
[243] F.R. Menter, M. Kuntz, R. Bender, A scale-adaptive simulation model for turbulent flow predictions, AIAA Paper, 2003.
[244] S.S. Girimaji, Partially-averaged Navier–Stokes model for turbulence: a reynolds-averaged Navier–Stokes to direct numerical simulation bridging
method, J. Appl. Mech. 73 (2006) 413–421.
[245] J. Rotta, Recent attempts to develop a generally applicable calculation method for turbulent shear flow layers, in: Proceedings AGARD Conference on
Turbulent Shear Flows, London, 1971.
[246] F.R. Menter, Y. Egorov. A scale-adaptive simulation model using two equation models. AIAA paper 2005-1095, Reno, NV, 2005.
[247] B. Basara, S. Krajnovic, Z. Pavlovic, P. Riugqvist, Performance analysis of partially-averaged Navier–Stokes methods for complex turbulent flows, in:
6th AIAA theoretical fluid mechanics conference, Honolulu, Hawaii, 2011.
C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732 731

[248] J. Ma, S.H. Peng, L. Davidson, F. Wang, A low Reynolds number variant of partially-averaged Navier–Stokes model for turbulence, Int. J. Heat Fluid
Flow 32 (2011) 652–669.
[249] L. Davidson, S.H. Peng, Embedded large-eddy simulation using the partially averaged Navier Stokes model, AIAA J. 51 (2013) 1066–1079.
[250] L. Davidson, Evaluation of the SST-SAS model: Channel flow, asymmetric diffuser and axi-symmetric hill, in: ECCOMAS CFD 2006, The Netherlands,
2006.
[251] B. Chaouat, R. Schiestel, A new partially integrated transport model for subgrid-scale stresses and dissipation rate for turbulent developing flows,
Phys. Fluids 17 (2005) 065106.
[252] R. Schiestel, A. Dejoan, Towards a new partially integrated transport model for coarse grid and unsteady turbulent flow simulations, Theor. Comput.
Fluid Dyn. 18 (2005) 443–468.
[253] S.T. Johansen, J. Wu, W. Shyy, Filter-based unsteady RANS computations, Int. J. Heat Fluid Flow 25 (2004) 10–21.
[254] N.S. Liu, T.H. Shih, Turbulence modelling for very large-eddy simulation, AIAA J. 44 (2006) 687–697.
[255] T.H. Shih, N.S. Liu, Modelling of internal reacting flows and external static stall flows using RANS and PRNS, Flow Turbul. Combust. 81 (2008) 279–
299.
[256] P. Batten, U. Goldberg, S. Chakravarthy, Interfacing statistical turbulence closures with large-eddy simulation, AIAA J. 42 (2004) 485–492.
[257] A. Ruprecht, T. Helmrich, I. Buntic, Very large eddy simulation for the prediction of unsteady vortex motion, in: Modelling Fluid Flow, Springer-Berlin
Heidelberg, 2004, pp. 229–246.
[258] J.B. Perot, J. Gadebusch, A self-adapting turbulence model for flow simulation at any mesh resolution, Phys. Fluids 19 (2007) 115105.
[259] J.B. Perot, J. Gadebusch, A stress transport equation model for simulating turbulence at any mesh resolution, Theor. Comput. Fluid Dyn. 23 (2009)
271–286.
[260] K.J. Hsieh, F.S. Lien, E. Yee, Towards a unified turbulence simulation approach for wall-bounded flows, Flow Turbul. Combust. 84 (2010) 193–218.
[261] D. Lakehal, M. Labois, A new modelling strategy for phase-change heat transfer in turbulent interfacial two-phase flow, Int. J. Multiphase Flow 37
(2011) 627–639.
[262] X. Han, S. Krajnović, An efficient very large eddy simulation model for simulation of turbulent flow, Int. J. Numer. Methods Fluids 71 (2013) 1341–
1360.
[263] P.R. Spalart, W-H Jou, M. Strelets, S.R. Allmaras, Comments on the feasibility of LES for wings, and on a hybrid RANS/LES approach, in: C. Lin, Z. Lin
(Eds.), Advances in DNS/LES, Creyden Press, Columbus, OH, 1997, pp. 137–147.
[264] P.R. Spalart, Detached-eddy simulation, Annu. Rev. Fluid Mech. 41 (2009) 181–202.
[265] A. Travin, M. Shur, M. Strelets, P. Spalart, Detached-eddy simulations past a circular cylinder, Flow Turbul. Combust. 63 (2000) 293–313.
[266] M.L. Shur, P.R. Spalart, M.K. Strelets, A.K. Travin, A hybrid RANS-LES approach with delayed-DES and wall-modelled LES capabilities, Int. J. Heat Fluid
Flow 29 (2008) 1638–1649.
[267] P. Spalart, S. Deck, M. Shur, K. Squires, M. Strelets, A. Travin, A new version of detached-eddy simulation, resistant to ambiguous grid densities, Theor.
Comput. Fluid Dyn. 20 (2006) 181–195.
[268] M. Strelets, Detached eddy simulation of massively separated flows, AIAA J. (2001) 1–18.
[269] M. Gritskevich, A. Garbaruk, J. Schiitze, F. Menter, Development of DDES and IDDES formulations for the k-x shear stress transport model, Flow Turb.
Comb. 88 (2012) 431–449.
[270] M. Leschziner, N. Li, F. Tessicini, Simulating flow separation from continuous surfaces: routes to overcoming the Reynolds number barrier, Philos.
Trans. Roy. Soc. A Math. Phys. Eng. Sci. 367 (2009) 2885–2903.
[271] L. Davidson, Hybrid LES-RANS: back scatter from a scale-similarity model used as forcing, Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 367 (2009)
2905–2915.
[272] F. Hamba, A hybrid RANS/LES simulation of turbulent channel flow, Theor. Comput. Fluid Dyn. 16 (2003) 387–403.
[273] L. Temmerman, M. Hadziabdic, M.A. Leschziner, K. Hanjalic, A hybrid two-layer URANS-LES approach for large eddy simulation at high Reynolds
numbers, Int. J. Heat Fluid Flow 26 (2005) 173–190.
[274] P. Tucker, L. Davidson, Zonal k–l based large eddy simulation, Comput. Fluids 33 (2004) 267–287.
[275] L. Davidson, S. Dahlstrom, Hybrid LES-RANS: an approach to make LES applicable at high Reynolds number, Int. J. Comput. Fluid Dyn. 19 (2005) 415–
427.
[276] X. Xiao, J.R. Edwards, H.A. Hassan, Blending functions in hybrid large eddy/Reynolds-averaged Navier–Stokes simulations, AIAA J. 42 (2004) 2508–
2515.
[277] I. Marusic, B.J. McKeon, P.A. Monkewitz, H.M. Nagib, A.J. Smits, K.R. Sreenivasan, Wall-bounded turbulent flows at high Reynolds numbers: recent
advances and key issues, Phys. Fluids 22 (2010) 065103.
[278] T. Mullin, Experimental studies of transition to turbulence in a pipe, Annu. Rev. Fluid Mech. 43 (2011) 1–24.
[279] N.V. Nikitin, Direct numerical modelling of three-dimensional turbulent flows in pipes of circular cross section, Fluid Dyn. 29 (1994) 749–758.
[280] J.G.M. Eggels, F. Unger, M.H. Weiss, J. Westerweel, R.J. Adrian, R. Friedrich, F.T.M. Nieuwstadt, Fully developed turbulent pipe flow: a comparison
between direct numerical simulation and experiment, J. Fluid Mech. 268 (1994) 175–210.
[281] Y. Zhang, A. Chandi, A.G. Tomboulides, S.A Orszag, Simulation of pipe flow, in: AGARD Conference Proceedings on Applications of Direct and Large
Eddy Simulations to Transition and Turbulence, 1994.
[282] P. Loulou, R.D. Moser, N.N. Mansour, B.J. Cantwell, Direct numerical simulation of incompressible pipe flow using a B-spline spectral method, NASA
report, 1997, pp. 1–138.
[283] A. Leonard, A. Wray, A new numerical method for simulation of three-dimensional flow in a pipe, Lect. Notes Phys. 170 (1982) 335–342.
[284] P. Orlandi, M. Fatica, Direct simulations of turbulent flow in a pipe rotating about its axis, J. Fluid Mech. 343 (1997) 43–72.
[285] P. Orlandi, D. Ebstein, Turbulent budgets in rotating pipes by DNS, Int. J. Heat Fluid Flow 21 (2000) 499–505.
[286] S. Schmidt, D.M. McIver, H.M. Blackburn, M. Rudman, G.J. Nathan, Spectral element based simulation of turbulent pipe flow, in: 14th Australasian
fluid mechanics conference, Adelaide University, Adelaide, Australia, 2001.
[287] J.M.J. Den Toonder, F.T.M. Nieuwstadt, Reynolds number effects in a turbulent pipe flow for low to moderate Re, Phys. Fluids 9 (1997) 3398–3409.
[288] C. Wagner, T.J. Huttl, R. Friedrich, Low-Reynolds-number effects derived from direct numerical simulations of turbulent pipe flow, Comput. Fluids 30
(2001) 581–590.
[289] K. Fukagata, N. Kasagi, Highly energy-conservative finite difference method for the cylindrical coordinate system, J. Comput. Phys. 181 (2002) 478–
498.
[290] M.P.B. Veenman, Statistical analysis of turbulent pipe flow: numerical approach, Faculty of Mechanical Engineering, Eindhoven University of
Technology, Eindhoven, 2004.
[291] N. Nikitin, A. Yakhot, Direct numerical simulation of turbulent flow in elliptical ducts, J. Fluid Mech. 532 (2005) 141–164.
[292] J. Kim, D. Kim, H. Choi, An immersed-boundary finite-volume method for simulations of flow in complex geometries, J. Comput. Phys. 171 (2001)
132–150.
[293] T.V. Voronova, N.V. Nikitin, Direct numerical simulation of the turbulent flow in an elliptical pipe, Comp. Math. Math. Phys. 46 (2006) 1378–1386.
[294] T.V. Voronova, N.V. Nikitin, Results of direct numerical simulation of turbulent flow in a pipe of elliptical cross-section, Fluid Dyn. 42 (2007) 201–211.
[295] X. Wu, P. Moin, A direct numerical simulation study on the mean velocity characteristics in turbulent pipe flow, J. Fluid Mech. 608 (2008) 81–112.
[296] B.J. Boersma, Direct numerical simulation of turbulent pipe flow up to a Reynolds number of 61000, J. Phys. Conf. Ser. 318 (2011) 042045.
[297] X. Wu, J.R. Baltzer, R.J. Adrian, Direct numerical simulation of a 30 R long turbulent pipe flow at R+ = 685: large- and very large-scale motions, J. Fluid
Mech. 698 (2012) 235–281.
732 C.D. Argyropoulos, N.C. Markatos / Applied Mathematical Modelling 39 (2015) 693–732

[298] G.K. Khoury, P. Schlatter, A. Noorani, P. Fischer, G. Brethouwer, A. Johansson, Direct numerical simulation of turbulent pipe flow at moderately high
Reynolds numbers, Flow Turbul. Combust. 91 (2013) 475–495.
[299] K. Lam, S. Banerjee, On the condition of streak formation in a bounded turbulent flow, Phys. Fluids A 4 (1992) 306–320.
[300] S. Komori, R. Nagaosa, Y. Murakami, S. Chiba, K. Ishii, K. Kuwahara, Direct numerical simulation of three-dimensional open-channel flow with zero-
shear gas-liquid interface, Phys. Fluids A Fluid Dyn. 5 (1993) 115–125.
[301] P. Lombardi, V.D. Angelis, S. Banerjee, Direct numerical simulation of near-interface turbulence in coupled gas–liquid flow, Phys. Fluids 8 (1996)
1643–1665.
[302] V.D. Angelis, Numerical investigation and modelling of mass transfer processes at shared gas–liquid interface (Ph.D. thesis), USCB, 1998.
[303] V.D. Angelis, P. Lombardi, P. Andreussi, S. Banerjee, Microphysics of scalar transfer at air-water interfaces, in: Proceedings of the IMA conference:
wind-over-wave coupling, perspectives and prospects, 1999, pp. 641–651.
[304] G.H. Yeoh, J. Tu, Computational Techniques for Multiphase Flows, A Butterworth-Heinemann, 2009.
[305] M. Fulgosi, D. Lakehal, S. Banerjee, V. De Angelis, Direct numerical simulation of turbulence in a sheared air–water flow with a deformable interface, J.
Fluid Mech. 482 (2003) 319–345.
[306] S. Banerjee, D. Lakehal, M. Fulgosi, Surface divergence models for scalar exchange between turbulent streams, Int. J. Multiphase Flow 30 (2004) 963–
977.
[307] M.Y. Lin, C.H. Moeng, W.G. Tsai, P.P. Sullivan, S.E. Belcher, Direct numerical simulation of wind-wave generation processes, J. Fluid Mech. 616 (2008)
1–30.
[308] D. Lakehal, M. Fulgosi, S. Banerjee, G. Yadigaroglu, Turbulence and heat exchange in condensing vapor–liquid flow, Phys. Fluids 20 (2008). 065101–18.
[309] P. Trontin, S. Vincent, J.L. Estivalezes, J.P. Caltagirone, Direct numerical simulation of a freely decaying turbulent interfacial flow, Int. J. Multiphase
Flow 36 (2010) 891–907.
[310] F. Unger, R. Friedrich, Large eddy simulation of fully-developed turbulent pipe flow, in: Proceedings 8th symposium on turbulent shear flows, 1991,
pp. 19-13-11–19-13-16.
[311] J.G.M. Eggels, F.T.M. Nieuwstadt, Large-eddy simulations of turbulent flow in an axially rotating pipe, in: Proceedings of the 9th symposium on
turbulent shear flows, 1993, pp. 310–313.
[312] B.J. Boersma, F.T.M. Nieuwstadt, Large-eddy simulation of turbulent flow in a curved pipe, J. Fluids Eng. 118 (1996) 248–254.
[313] Z. Yang, Large eddy simulation of fully developed turbulent flow in a rotating pipe, Int. J. Numer. Methods Fluids 33 (2000) 681–694.
[314] M. Rudman, H.M. Blackburn, Large eddy simulation of turbulent pipe flow, in: Second International Conference on CFD in the Minerals and Process
Industries, Melbourne, Australia, 1999, pp. 503–508.
[315] A.A. Feiz, M. Ould-Rouis, G. Lauriat, Large eddy simulation of turbulent flow in a rotating pipe, Int. J. Heat Fluid Flow 24 (2003) 412–420.
[316] S.A. Jordan, The turbulent character and pressure loss produced by periodic symmetric ribs in a circular duct, Int. J. Heat Fluid Flow 24 (2003) 795–
806.
[317] S. Vijiapuraru, J. Cui, Large eddy simulation of fully developed turbulent pipe flow, ASME Conf. Proc. (2004) 675–679.
[318] S. Vijiapuraru, J. Cui, Performance of turbulence models for flows through rough pipes, App. Math. Modell. 34 (2010) 1458–1466.
[319] S.Y. Jung, Y.M. Chung, Large-eddy simulation of accelerated turbulent flow in a circular pipe, Int. J. Heat Fluid Flow 33 (2012) 1–8.
[320] D. Lakehal, LEIS for the prediction of turbulent multifluid flows applied to thermal-hydraulics applications, Nucl. Eng. Des. 240 (2010) 2096–2106.
[321] S. Reboux, P. Sagaut, D. Lakehal, Large-eddy simulation of sheared interfacial flow, Phys. Fluids 18 (2006) 105105.
[322] E.D. Christensen, Large eddy simulation of spilling and plunging breakers, Coast. Eng. 53 (2006) 463–485.
[323] P. Lubin, S. Vincent, S. Abadie, J.-P. Caltagirone, Three-dimensional large eddy simulation of air entrainment under plunging breaking waves, Coast.
Eng. 53 (2006) 631–655.
[324] D. Lakehal, P. Liovic, Turbulence structure and interaction with steep breaking waves, J. Fluid Mech. 674 (2011) 522–577.
[325] P. Liovic, D. Lakehal, Multi-physics treatment in the vicinity of arbitrarily deformable gas–liquid interfaces, J. Comput. Phys. 222 (2007) 504–535.
[326] F.C.K. Ting, J.T. Kirby, Dynamics of surf-zone turbulence in a spilling breaker, Coast. Eng. 27 (1996) 131–160.
[327] P.J. Roache, Quantification of uncertainty in computational fluid dynamics, Annu. Rev. Fluid Mech. 29 (1997) 123–160.
[328] W.L. Oberkampf, T.G. Trucano, Verification and validation in computational fluid dynamics, Prog. Aerosp. Sci. 38 (2002) 209–272.
[329] C.J. Roy, Review of code and solution verification procedures for computational simulation, J. Comput. Phys. 205 (2005) 131–156.
[330] F. Stern, R. Wilson, J. Shao, Quantitative V&V of CFD simulations and certification of CFD codes, Int. J. Numer. Methods Fluids 50 (2006) 1335–1355.
[331] ASME, Standards for Verification and Validation in Computational Fluid Dynamics and Heat Transfer, The American Society of Mechanical Engineers
(ASME), 2009.
[332] NEA, Best practice guidelines for the use of CFD in nuclear reactor safety applications, Nuclear Energy Agency (NEA), Committee on the safety of
nuclear installations, NEA/CSNI/R(2007) 5, 2007.
[333] ERCOFTAC, Best Practice Guidelines, European Research Community on Flow, Turbulence and Combustion (ERCOFTAC), Special Interest Group on
Quality and Trust in Industrial CFD, Ver. 1, 2000.
[334] AIAA, Guide for the Verification and Validation of Computational Fluid Dynamics Simulations, AIAA Guide G-077-1988, 1998.
[335] M.K. Patel, M. Cross, N.C. Markatos, An assessment of flow-oriented schemes for reducing false diffusion, Int. J. Numer. Methods Eng. 26 (1988) 2279–
2304.
[336] M.K. Patel, N.C. Markatos, M. Cross, Methods of reducing false-diffusion errors in convection-diffusion problems, Appl. Math. Modell. 9 (1985) 302–
306.
[337] M.K. Patel, N.C. Markatos, An evaluation of eight discretization schemes for two-dimensional convection-diffusion equations, Int. J. Numer. Methods
Fluids 6 (1986) 103–112.
[338] M.K. Patel, M. Cross, A.C.H. Mace, N.C. Markatos, An evaluation of eleven discretization schemes for predicting elliptic flow and heat transfer in
supersonic jets, Int. J. Heat Mass Transfer 30 (1987) 1907–1925.
[339] D.P. Karadimou, N.C. Markatos, A novel flow-oriented discretization scheme for reducing false-diffusion in three-dimensional flows: an application in
the indoor environment, Atmos. Environ. 61 (2012) 327–339.

You might also like