Computational Fluid Dynamics: Incompressible Turbulent Flows 1st Edition Takeo Kajishima
Computational Fluid Dynamics: Incompressible Turbulent Flows 1st Edition Takeo Kajishima
com
https://textbookfull.com/product/computational-fluid-
dynamics-incompressible-turbulent-flows-1st-edition-takeo-
kajishima/
OR CLICK HERE
DOWLOAD EBOOK
https://textbookfull.com/product/computational-fluid-dynamics-for-
incompressible-flow-first-edition-roychowdhury/
textbookfull.com
https://textbookfull.com/product/computational-methods-for-fluid-
dynamics-joel-h-ferziger/
textbookfull.com
https://textbookfull.com/product/computational-fluid-dynamics-in-food-
processing-second-edition-edition-sun/
textbookfull.com
https://textbookfull.com/product/computational-fluid-dynamics-a-
practical-approach-3rd-edition-jiyuan-tu/
textbookfull.com
Uncertainty Quantification in Computational Fluid Dynamics
and Aircraft Engines 1st Edition Francesco Montomoli
https://textbookfull.com/product/uncertainty-quantification-in-
computational-fluid-dynamics-and-aircraft-engines-1st-edition-
francesco-montomoli/
textbookfull.com
https://textbookfull.com/product/computational-fluid-dynamics-for-
built-and-natural-environments-zhiqiang-john-zhai/
textbookfull.com
https://textbookfull.com/product/computational-fluid-dynamics-in-food-
processing-2nd-edition-da-wen-sun-editor/
textbookfull.com
https://textbookfull.com/product/cellular-flows-topological-
metamorphoses-in-fluid-mechanics-1st-edition-vladimir-shtern/
textbookfull.com
https://textbookfull.com/product/non-newtonian-fluid-mechanics-and-
complex-flows-angiolo-farina/
textbookfull.com
Takeo Kajishima · Kunihiko Taira
Computational
Fluid Dynamics
Incompressible Turbulent Flows
Computational Fluid Dynamics
Takeo Kajishima Kunihiko Taira
•
Computational Fluid
Dynamics
Incompressible Turbulent Flows
123
Takeo Kajishima Kunihiko Taira
Department of Mechanical Engineering Florida State University
Osaka University Tallahassee, FL
Osaka USA
Japan
The majority of flows encountered in nature and engineering applications are tur-
bulent. Although turbulence has been studied as part of classical mechanics for over
a century, it is still one of the unsolved problems in physics and remains to be an
active area of research. The field of study that uses numerical simulation to examine
fluid dynamics is called Computational Fluids Dynamics (CFD). There is great
demand in utilizing CFD for analyzing problems ranging from turbulent flows
around aircraft and ground vehicles to larger scale problems related to weather
forecasting and environmental assessment. Such demands are very likely to grow in
the coming years as the engineering community at large pursues improvements in
energy efficiency and performance for various fluid-based systems.
Presently, there are several commercial CFD solvers released with turbulence
analysis capability. With software creating beautiful visualizations of turbulent
flows, it may appear that any type of flows can be numerically predicted. While
there may be some truth to such capability, it is still difficult to solve most turbulent
flow problems without relying on companion experiments. That raises a question of
why we still are not able to perfectly predict the behavior of turbulent flows. The
dynamics of turbulent flows obeys the Navier–Stokes equations, upon which CFD
solvers are based. However, turbulence exhibits flow structures over a wide range
of spatial and temporal scales that all interacts amongst them in a complex nonlinear
manner. That means that the spatial grid must be fine enough to resolve the smallest
scales in turbulent flows while ensuring that the computational domain is large
enough to encompass the largest flow structures. Such grid requirement becomes
increasingly costly as we tackle flows at a higher Reynolds number. Despite the
significant improvement in the computational capability with recent
high-performance computers, we still do not expect computers to be able to handle
these large grids for very high Reynolds number flows.
For this particular reason, turbulence is not likely to be completely solved in the
near future. Flow physics taking place at scales below the resolvable scales must be
represented with appropriate models, referred to as turbulence models. Presently,
there is not a universally accepted turbulence model or numerical algorithm that can
yield a solution unaffected by discretizations of the flow field. Thus, CFD should
vii
viii Preface
continue to be an active field of research with efforts focused towards predicting the
essential features of turbulent flows with turbulence models. As such, engineers and
scientists using CFD must understand how the governing equations are numerically
solved. We must also be equipped with the ability to correctly interpret the
numerical solution. With these points in mind, we should construct a necessary and
sufficient computer program appropriate to simulate the fluid flow of interest. For
commercial software, sufficient details on the solver technique should be provided
in the reference manual so that users can determine whether the solver can be
appropriately used for the problem at hand.
This book describes the fundamental numerical methods and approaches used to
perform numerical simulations of turbulent flows. The materials presented herein
are aimed to provide the basis to accurately analyze unsteady turbulent flows. This
textbook is intended for upper level undergraduate and graduate students who are
interested in learning CFD. This book can also serve as a reference when devel-
oping incompressible flow solvers for those already active in CFD research. It is
assumed that readers have some knowledge of fluid mechanics and partial differ-
ential equations. This textbook does not assume the readers to have advanced
knowledge of numerical analysis.
This textbook aims to enable readers to construct his or her own CFD code from
scratch. The present textbook covers the numerical methods required for CFD and
places emphasis on the incompressible flow solver with detailed discussions on
discretization techniques, boundary conditions, and turbulent flow physics. The
introduction to CFD and the governing equations are offered in Chap. 1, followed
by the coverage of basic numerical methods in Chap. 2. Incompressible flow sol-
vers are derived and discussed in detail in Chaps. 3 and 4. We also provide dis-
cussions on the immersed boundary methods in Chap. 5. A brief overview on
turbulent flows is given in Chap. 6 with details needed for analyzing turbulent flows
using Reynolds-Averaged Navier–Stokes equations (RANS) and Large-Eddy
Simulation (LES) provided in Chaps. 7 and 8, respectively. At the end of the
book, an appendix is attached to offer details on the generalized coordinate system,
Fourier analysis, and modal decomposition methods.
A large portion of the present book is based on the material taught over the years
by the first author for the course entitled “Computational Fluid Dynamics and
Turbulent Flows” at Osaka University and his textbook entitled, “Numerical
Simulations of Turbulent Flows” (1st and 2nd editions in Japanese) that has been
available in Japan since 1999. Chapters 1–4 and 6–8 as well as Appendices A and B
in the present book are founded heavily on the Japanese version by Kajishima. The
present textbook enjoys additions of stability analysis (Sect. 2.5), immersed
boundary methods (Chap. 5), and modal decomposition methods (Sect. 6.3.7 and
Appendix C) by Taira based on the courses taught at the Florida State University.
Furthermore, exercises have been added after each chapter to provide supplemental
materials for the readers.
The preparation of this book has benefited greatly from comments, feedback,
and encouragements from Takashi Ohta, Shintaro Takeuchi, Takeshi Omori, Yohei
Morinishi, Shinnosuke Obi, Hiromochi Kobayashi, Tim Colonius, Clarence
Preface ix
xi
xii Contents
1.1 Introduction
Numerical simulations, along with experiments and theoretical analysis, are often
used as a tool to support research and development in science and engineering.
The use of simulations has been popularized by the development and wide-spread
availability of computers. Since numerical computations are advantageous to exper-
iments from the aspects of speed, safety, and cost in many cases, their uses have been
widely accepted in the industry. Simulations have also become a valuable tool in
fundamental research due to its ability to analyze complex phenomena that may be
difficult to study with experimental measurements or theoretical analysis. Reflecting
upon these trends, the adjective computational is now widely used to describe sub-
fields that utilize simulation in various disciplines, such as computational physics
and computational chemistry.
The field of study concerned with analyzing various types of fluid flows with
numerical simulations and developing suitable simulation algorithms is known as
computational fluid dynamics (CFD). Applications of CFD can be found in the analy-
sis of the following studies but not limited to
• Flows around aircraft, ships, trains, and automobiles;
• Flows in turbo-machineries;
• Biomedical and biological flows;
• Environmental flows, civil engineering, and architecture;
• Large-scale flows in astrodynamics, weather forecasting, and oceanography.
The flows in these settings usually do not have analytical expressions because of
the complex physics arising from boundary geometry, external forcing, and fluid
properties. In CFD, flow physics is analyzed and predicted by numerically solving
the governing equations and reproducing the flow field with the use of computers.
In general, a fluid can be viewed as a continuum, for which there are established
conservation laws for mass, momentum, and energy. For many fluids that are used in
engineering applications, there are well-accepted constitutive relations. The equation
of state for gases or equations for phase transformation or chemical reactions are also
included in the system of equations as needed. The objective of flow simulations is
to numerically solve such system of equations with appropriate initial and boundary
conditions to replicate the actual flow. In the simulation, the flow field is represented
by variables such as velocity, pressure, density, and temperature at discrete set of
points. The evolution of these variables over time is tracked to represent the flow
physics.
There is always the question of whether a simulation correctly reproduces the
flow physics, because we are representing the continuum on discrete points. We
can only obtain reliable numerical solutions when the simulation methodology is
validated against experimental measurements or theoretical solutions. The use of such
validated method must be limited to parameters that are within the applicable range.
While there has been numerous achievements with computational fluid dynamics in
the design fields and fundamental research, the development of accurate simulation
methods that are widely applicable and more robust will continue in the future.
Numerical simulation of fluid flow follows the steps laid out in Fig. 1.1.
Thus, to perform numerical simulations of fluid flows, it is not sufficient to only have
the understanding of fluid mechanics. One must have knowledge of numerical analy-
sis for discretization schemes and numerical algorithms, as well as computer science
for programing and visualization. It goes without saying that the combination of
these fields is necessary for CFD. To ensure that the results obtained from numerical
simulation of fluid flow are reliable, we must also be concerned with verification and
validation. We will discuss this in further details in Sect. 1.6.
1.2 Overview of Fluid Flow Simulations 3
experiments
(dye visualization)
Approximation
Physical model
Governing equations
(partial differential equations)
Discretization
Grid generation
Numerical methods
Code development
Data visualization/analysis
simulation 3D printing
(vorticity contour)
Fig. 1.1 General process of simulating fluid flows. Inserted visualizations are for unsteady flow
over a pitching plate (Reprinted with permission from [8]; copyright 2014, AIP Publishing LLC)
4 1 Numerical Simulation of Fluid Flows
In this section, we present the governing equations of fluid flows. There are two
formulations one can use to describe a flow field: the Eulerian and Lagrangian rep-
resentations. The Eulerian representation describes the flow field with functions of
space and time. The Lagrangian representation on the other hand describes the flow
field following individual fluid elements in the flow. In this book, we discretize the
governing equations on a grid which is based on the Eulerian formulation. For that
reason, the discussions herein are based on the Eulerian representation of the flow
field.
For details on the two representations of a flow field and the derivation of the
governing equations starting from vector and tensor analysis, we list Currie [4],
Panton [15], and Aris [1] as references.
The governing equations for fluid flows consist of the conservation laws for mass,
momentum, and energy.
First, let us consider mass conservation. Denoting the density (mass per unit
volume) of the fluid by ρ, we perform a budget analysis for a control volume. We let
the volume and surface of the control volume be V and S, respectively, with the unit
normal vector n on the surface (directed outward) and the flow velocity u, as shown
in Fig. 1.2. The time rate of change of the mass within the control volume consists of
the mass flux going in and out of the volume (ρu) · n through the surface assuming
there is no mass source or sink:
∂ρ
dV = − (ρu) · ndS. (1.1)
V ∂t S
1.3 Governing Equations of Fluid Flows 5
Here, mass influx is negative (ρu · n < 0) and efflux is positive (ρu · n > 0).
Using Gauss’ theorem, the above equation can be written only with a volume
integral
∂ρ
+ ∇ · (ρu) dV = 0. (1.2)
V ∂t
Equations (1.1) and (1.2) are integral representations of mass conservation. Since
Eq. (1.2) must hold for any arbitrary control volume, the integrand should be zero.
Thus, we have
∂ρ
+ ∇ · (ρu) = 0, (1.3)
∂t
which is the differential form of mass conservation.
Following similar control volume analysis, the equations for momentum and
energy conservation can be derived. We can represent these mass, momentum, and
energy conservation laws all together in a single equation using tensors. The integral
representation of the three conservation laws for a control volume V becomes
∂Λ
dV = − Π · ndS + Γ dV (1.4)
∂t
V S
V
∂Λ
+ ∇ · Π − Γ dV = 0 (1.5)
V ∂t
∂Λ
+∇ ·Π = Γ. (1.6)
∂t
Here, the vector Λ denotes the conserved quantities (per unit volume)
⎡ ⎤
ρ
Λ = ⎣ ρu ⎦ , (1.7)
ρE
where E is the total energy per unit mass, which is comprised of the internal energy
(per unit mass) e and the kinetic energy k
6 1 Numerical Simulation of Fluid Flows
E = e + k. (1.8)
The term ρu can be recognized as the momentum per unit volume in the second row
of Eq. (1.7) and also as the mass flux vector for the first row of Eq. (1.10). The term
ρuu − T is the momentum flux tensor in the momentum equation which consists
of ρuu that describes the flux of momentum ρu moving at a velocity u and T that
represents the momentum exchange due to the stress at the surface of the control
volume. In the energy conservation equation, ρE u is the flux of energy, T · u is the
work performed by stress, and q is the heat flux. By taking an inner product of Π
and the unit normal vector n, we obtain the physical quantity that passes through the
control surface per unit time and area. Since n represents the outward unit normal
vector on a control surface S in Eq. (1.4), positive and negative Π · n correspond to
efflux and influx, respectively, from the control surface.
In what follows, we assume that there is no sink, source, or heat generation within
the volume of interest. If a body force f acts on the fluid, there is production of
momentum and work performed by the force, making the right-hand side of Eq. (1.6)
become ⎡ ⎤
0
Γ = ⎣ ρ f ⎦. (1.11)
ρu · f
Note that Eqs. (1.4) and (1.5) represent the change in conserved variables that is
attributed to the flux across surface of the control volume and the source within the
volume.
In order to solve the governing equations for fluid flow, we need to match the number
of unknowns to the number of equations. This is referred to as the closure of the
system of equations. For the flow equations, we need to express the stress T and heat
flux q in the flux term Π by ρ, u, and E. These relations are called the constitutive
equations.
1.3 Governing Equations of Fluid Flows 7
For a Newtonian fluid, the stress tensor T is expressed as the sum of pressure and
viscous stress in the following manner
1
T = − p I + 2μ D − I∇ · u , (1.12)
3
in which the Stokes relation has been used. Here, I is the identity tensor, p is the
static pressure, μ is the dynamic viscosity, and D is the rate-of-strain tensor given
by
1
D= (∇u)T + ∇u . (1.13)
2
For the heat flux q, we can use the Fourier’s law
q = −k∇T, (1.14)
where T is the absolute temperature and k is the thermal conductivity. Note that μ
and k can be expressed as functions of T for Newtonian fluids (i.e., μ(T ) and k(T )).
We have introduced T and p, which can be related to ρ by the equation of state.
Here, we make an assumption that the fluid is an ideal gas in thermodynamic equi-
librium. The ideal gas law states that
where γ = c p /cv is the specific heat ratio, cv is the heat capacity at constant volume,
c p (= cv + R) is the heat capacity at constant pressure, and R is the gas constant.
The internal energy e is
e = cv T (1.16)
The mass conservation equation is also referred to as the continuity equation and is
∂ρ
+ ∇ · (ρu) = 0 (1.17)
∂t
for flows without sinks or sources. We can consider the time rate of change and
transport of some physical quantity φ by the velocity field u to be decomposed as
8 1 Numerical Simulation of Fluid Flows
∂(ρφ) ∂φ ∂ρ
+ ∇ · (ρuφ) = ρ + u · ∇φ + φ + ∇ · (ρu) , (1.18)
∂t ∂t ∂t
where the second term on the right-hand side becomes zero due to continuity,
Eq. (1.17). Accordingly, we have
∂(ρφ) Dφ ∂φ
+ ∇ · (ρuφ) = ρ =ρ + u · ∇φ , (1.19)
∂t Dt ∂t
1 Advective form is often called convective form. We however use the term advective form to be
consistent with the use of the term advection instead of convection (= advection + diffusion) for
preciseness.
1.3 Governing Equations of Fluid Flows 9
corresponds to the acceleration of a fluid element, we can notice that Eq. (1.22)
describes Newton’s second law (mass × acceleration = force) per unit mass.
Taking an inner product of the momentum equation, Eq. (1.22), with the velocity
u, we arrive at the conservation of kinetic energy
Dk
ρ = u · (∇ · T ) + ρu · f . (1.24)
Dt
Subtracting the above equation from the conservation of total energy
DE
ρ = ∇ · (T · u) − ∇ · q + ρu · f , (1.25)
Dt
we obtain the conservation of internal energy
De
ρ = T : (∇u) − ∇ · q, (1.26)
Dt
where T : S denotes the contraction of tensors (i.e., Ti j S ji ). The set of equations
consisting of the mass, momentum, and energy equations for a Newtonian fluid is
called the Navier–Stokes equations.
Up to this point, we have used vector notation in the governing equations for fluid
flows. We can also utilize what is called indicial notation to represent the components
for a Cartesian coordinate system, in which we denote the coordinates with x1 = x,
x2 = y, and x3 = z and the corresponding velocity components with u 1 = u, u 2 = u,
and u 3 = w.
We then can express the mass conservation as
∂ρ ∂(ρu j )
+ = 0. (1.27)
∂t ∂x j
∂(ρE) ∂
+ (ρEu j − Ti j u i + q j ) = ρu i f i (1.30)
∂t ∂x j
When the same index appears twice in the same term, summation over that index is
implied (summation convention). That is, in two dimensions, a j b j = 2j=1 a j b j , and
in three dimensions, a j b j = 3j=1 a j b j . When summation is not to be performed, it
will be noted in the text. The symbol for the index can be different in the summation
but results in the same sum (i.e., a j b j = ak bk ). The symbol δi j appears often when
using indicial notation and is called the Kronecker delta, defined as
1 i= j
δi j = (1.34)
0 i = j
This is the component-wise representation of the basis tensor I for the Cartesian
coordinate system. Note that the trace in three dimensions is δkk = 3 (contraction of
δi j ).
Dρ ∂ρ
= + u · ∇ρ = 0. (1.35)
Dt ∂t
Note that incompressibility does not necessarily mean that the density ρ is constant.
This continuity equation becomes much more complex for multispecies system with
diffusion, even if the flow can be treated as incompressible [9]. With Eq. (1.35), the
continuity equation, Eq. (1.17), turns into
∇ · u = 0, (1.36)
T = − p I + 2μ D. (1.37)
If we can treat viscosity to be a constant, the above equation can be further simplified
to become
∂u ∇p
+ ∇ · (uu) = − + ν∇ 2 u + f , (1.39)
∂t ρ
∇2 p
= −∇ · ∇ · (uu) + ∇ · f , (1.40)
ρ
where ∇ 2 is the Laplacian operator. Here, the flow field is constrained by incom-
pressibility, Eq. (1.36), at all times and the corresponding pressure field is determined
from the instantaneous flow field. Equation (1.40) is called the pressure Poisson equa-
tion and is an elliptic partial differential equation that is solved as a boundary value
problem. While pressure for compressible flow is provided thermodynamically by
the equation of state, Eq. (1.15), the pressure for incompressible flow is not deter-
mined using thermodynamics. This leads to adopting different numerical methods for
incompressible and compressible flows. While both incompressible and compressible
12 1 Numerical Simulation of Fluid Flows
De
ρ = μ D : D − ∇ · q. (1.41)
Dt
With the use of Eqs. (1.14) and (1.16), the equation for the temperature field reads
DT
ρcv = μ D : D + k∇ 2 T. (1.42)
Dt
The first term on the right-hand side represents the heat generation due to fluid
friction (viscous effect). If the influence of friction on the temperature field is small,
the temperature equation simplifies to
DT
ρcv = k∇ 2 T. (1.43)
Dt
Under this assumption, we can consider the kinetic and internal energies separately
for incompressible flows. Since the conservation of kinetic energy depends only
passively on the conservation of mass and momentum, it is not necessary to explicitly
handle the kinetic energy in the fluid flow analysis. The conservation of internal
energy can be used as a governing equation for temperature. In the discussions to
follow, we do not consider the temperature equation.
Component-wise Representation
As a summary of the above discussion and for reference, let us list the governing
equations for incompressible flow with constant density and viscosity for a Cartesian
coordinate system using the component-wise representation. The continuity equation
(1.36) is
∂u i
=0 (1.44)
∂xi
∂u i ∂(u i u j ) ∂u i ∂u i 1 ∂p ∂2ui
+ = + uj =− +ν + fi . (1.45)
∂t ∂x j ∂t ∂x j ρ ∂xi ∂x j ∂x j
2 For incompressible flow, the Mach number M = u/a (the ratio of characteristic velocity u and
sonic speed a) is zero or very small. That means that the acoustic wave propagation is very fast
compared to the hydrodynamic wave propagation. In the limit of M → 0, the acoustic propagation
is considered to take place instantaneously over the whole domain, which leads to the appearance
of ellipticity in the governing equations.
1.3 Governing Equations of Fluid Flows 13
The pressure Poisson equation derived from the above two equations is
1 ∂2 p ∂ 2 (u i u j ) ∂ f i ∂u i ∂u j ∂ fi
=− + =− + . (1.46)
ρ ∂xi ∂xi ∂xi ∂x j ∂xi ∂x j ∂xi ∂xi
Note that Eqs. (1.45) and (1.46) make use of Eq. (1.44).
As we have seen above, the conservation laws that describe the flow are represented
by partial differential equations (PDEs). Hence, it is important to understand the
characteristics of the governing PDEs to numerically solve for the flow field. It
is customary to discuss the classification of PDEs and the theory of characteristic
curves, but unless readers go beyond the scope of this book or become involved in
the development of advanced numerical algorithms, there is not a critical need to
dive deeply into the theory of PDEs.
Depending on how information travels, the second-order PDEs are classified into
elliptic, parabolic, and hyperbolic PDEs as shown in Table 1.1. These types are not
influenced by the choice of coordinate systems but can be affected by the location
in a flow field. For example, let us consider flow over a bluff body. Flow away from
the body and outside of the boundary layer (without any vorticity) can be treated as
potential flow which is described by an elliptic PDE. However, in the region right
next to the surface of the body (boundary layer), viscous diffusion becomes the
dominant physics describing the flow which can be captured by a parabolic PDE.
Hence, the classification of PDEs is in general performed locally. We will let other
textbooks [10, 13] on PDEs describe the classification of PDEs, characteristic curves,
and initial/boundary value problems. For the purpose of this book, we will focus on
presenting the PDE types of the governing equations for flow in a brief fashion. The
classification of PDEs is based on the existence of a unidirectional coordinate [16].
The temporal axis t is obviously unidirectional (past to future). On the other hand,
the spatial coordinate xi is bidirectional.
For viscous incompressible flows in general, the PDEs for unsteady flow (depen-
dent on t) are parabolic and the PDEs for steady flow (independent of t) are elliptic.
In some steady cases where the influence from downstream can be neglected, the flow
may be solved by marching from upstream to downstream. This is called parabolic
approximation (parabolization). One can also solve for steady flow by leaving the
temporal derivative term and reframing the problem as an unsteady one. Once the
solution is converged to the steady profile after sufficient time advancement, the time
derivative term would vanish and the flow field becomes the solution to the original
elliptic PDE. Such approach could be described as a parabolic technique for solving
an elliptic PDE.
For Eulerian methods, variables such as velocity, pressure, and density are determined
at a large number of discrete points to represent the motion of a liquid or a gas as
a continuum. A polygon (in two-dimensional and polyhedron in three-dimensional
space) made from local collection of discrete points (vertices) is called a cell and
the space filled by these cells is referred to as a grid (or mesh). Physical variables of
interest are positioned at various locations on the cells chosen to satisfy particular
numerical properties.
A few representative Eulerian grids are shown in Fig. 1.3. For a Cartesian grid,
the governing equations for the flow takes the simplest form and makes spatial
discretization effortless. However, it is not suitable for discretizing a flow field around
a body of complex geometry, as illustrated in Fig. 1.3a. Even a body with rather simple
geometry, such as a sphere or a circular cylinder, would requires a very fine mesh
near the body boundary to resolve the flow. One solution to this issue is to employ an
immersed boundary method that generates a body without regard to the underlying
grid. Chapter 5 is devoted to the immersed boundary method. Another remedy to
this problem is to use a curvilinear coordinate grid that fits around the boundary, as
shown in Fig. 1.3b. Such curvilinear grid is called a boundary-fitted coordinate grid
or a body-fitted coordinate grid (BFC). A transformation of such BFC in physical
Fig. 1.3 Grids for fluid flow simulations (two-dimensional). a Cartesian grid. b Curvilinear grid.
c Unstructured grid
1.4 Grids for Simulating Fluid Flows 15
(a) (b)
A D
η
ξ Δη = 1
C D
B A η
y Δξ = 1
Computational domain
x B
ξ
C
Fig. 1.4 Mapping between the physical space and computation space for a boundary-fitted grid. a
Physical space. b Computational space
Fig. 1.6 Examples of hybrid grids. a Overset grid. b Patched grid. c Hybrid grid
cal boundary-fitted grids that are used to discretize the computational domain around
an airfoil. The O-grid places the grid around a body efficiently with low skewness
in general. For bodies with sharp corners (cusp), for example the trailing edge of an
airfoil, there can be highly skewed grids. The C-grid is aligned with the flow around
the body and is able to generate unskewed grids near the trailing edge. While such
arrangement of grids is suitable for viscous flows, there may be unnecessarily large
number of grids downstream of the wing. For airfoil cascade (arrangement of a series
of blades, such as in turbines) or flow over a body in a channel, multi-block approach
is often utilized with H-grids and L-grids (not shown).
It would be ideal to discretize the flow field with only one type of grid. With
the development of grid generation techniques, it has become possible to generate a
mesh of a single type even for somewhat complex geometries. However, generating
high-quality meshes remains a challenge (e.g., mesh with low skewness, mesh with
necessary and sufficient resolution). In many cases, one cannot know where the grid
should be refined a priori.
For flows with bodies of complex geometry, with multiple bodies, or with bodies
that move (relative to each other) or deform, we can consider the use of multiple
types of grids, as illustrated in Fig. 1.6. In most cases, information is transmitted
during calculation from one grid to the other at the overlap or along the interface of
the grids.
Concentrating grids in regions where the flow exhibits changes in its features
is effective for attaining accurate solutions. For example, we know a priori that the
boundary layer near the wall has large velocity gradients. Thus, it would be beneficial
to place a large number of grid points to resolve the flow there. One can also consider
adaptively generating additional grids at regions where the flow shows large gradients
in the variables of interest. For unstructured grids, this would be handled by simply
adding extra vertices in that region. Such approach is referred to as adaptive mesh
refinement and is often used to capture shock waves and flames. In case of flows with
time varying boundary shapes, such as the waves around a ship, a moving adaptive
mesh is used.
In what follows, we assume that the grid has been generated for the computational
domain. For details on grid generation, readers should consult with [6, 19].
Exploring the Variety of Random
Documents with Different Content
NOUVELLES DES ÎLES
INFORTUNÉES
A Jules Renard.
Updated editions will replace the previous one—the old editions will
be renamed.
1.D. The copyright laws of the place where you are located also
govern what you can do with this work. Copyright laws in most
countries are in a constant state of change. If you are outside the
United States, check the laws of your country in addition to the
terms of this agreement before downloading, copying, displaying,
performing, distributing or creating derivative works based on this
work or any other Project Gutenberg™ work. The Foundation makes
no representations concerning the copyright status of any work in
any country other than the United States.
1.E.6. You may convert to and distribute this work in any binary,
compressed, marked up, nonproprietary or proprietary form,
including any word processing or hypertext form. However, if you
provide access to or distribute copies of a Project Gutenberg™ work
in a format other than “Plain Vanilla ASCII” or other format used in
the official version posted on the official Project Gutenberg™ website
(www.gutenberg.org), you must, at no additional cost, fee or
expense to the user, provide a copy, a means of exporting a copy, or
a means of obtaining a copy upon request, of the work in its original
“Plain Vanilla ASCII” or other form. Any alternate format must
include the full Project Gutenberg™ License as specified in
paragraph 1.E.1.
• You pay a royalty fee of 20% of the gross profits you derive
from the use of Project Gutenberg™ works calculated using the
method you already use to calculate your applicable taxes. The
fee is owed to the owner of the Project Gutenberg™ trademark,
but he has agreed to donate royalties under this paragraph to
the Project Gutenberg Literary Archive Foundation. Royalty
payments must be paid within 60 days following each date on
which you prepare (or are legally required to prepare) your
periodic tax returns. Royalty payments should be clearly marked
as such and sent to the Project Gutenberg Literary Archive
Foundation at the address specified in Section 4, “Information
about donations to the Project Gutenberg Literary Archive
Foundation.”
• You comply with all other terms of this agreement for free
distribution of Project Gutenberg™ works.
1.F.
1.F.4. Except for the limited right of replacement or refund set forth
in paragraph 1.F.3, this work is provided to you ‘AS-IS’, WITH NO
OTHER WARRANTIES OF ANY KIND, EXPRESS OR IMPLIED,
INCLUDING BUT NOT LIMITED TO WARRANTIES OF
MERCHANTABILITY OR FITNESS FOR ANY PURPOSE.
Please check the Project Gutenberg web pages for current donation
methods and addresses. Donations are accepted in a number of
other ways including checks, online payments and credit card
donations. To donate, please visit: www.gutenberg.org/donate.
Most people start at our website which has the main PG search
facility: www.gutenberg.org.
Our website is not just a platform for buying books, but a bridge
connecting readers to the timeless values of culture and wisdom. With
an elegant, user-friendly interface and an intelligent search system,
we are committed to providing a quick and convenient shopping
experience. Additionally, our special promotions and home delivery
services ensure that you save time and fully enjoy the joy of reading.
textbookfull.com