You are on page 1of 227

Computational Fluid Dynamics

MMAE 517

Illinois Institute of Technology


c 2010 K. W. Cassel

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 1 / 454

Introduction to CFD

CFD Advertisements...

“CFD is a modern and rather special expression


of fluid mechanics. It is an endlessly fascinating,
interdisciplinary blend of basic numerical methods,
a solid understanding of both experimental
and anayltical fluid dynamics, software engineering
and pragmatism.”
W.N. Dawes (JFM 1996)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 2 / 454
Introduction to CFD

Advantages and disadvantages of analytical, computational and experimental


approaches to fluid mechanics:
Analytical:
+ Provides exact solutions to governing equations.
+ Gives physical insight, e.g. relative importance of different effects.
+ Can consider hypothetical problems (e.g. inviscid, incompressible,
zero-gravity, etc.).
− Exact solutions only available for simple problems and geometries.
− Of limited direct value in design.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 3 / 454

Introduction to CFD

Computational:
+ Address more complex problems (physics and geometries).
+ Can consider hypothetical flows ⇒ test theoretical models.
+ Provides detailed solutions ⇒ good understanding of flow, e.g. does
separation occur?
+ Perform parametric studies.
+ Can easily try different configurations, e.g. geometry, boundary conditions,
etc... ⇒ important in design.
+ Computers becoming faster and cheaper ⇒ range of CFD expanding.
+ Increased potential using parallel processing.
+ More cost effective and faster than experimental prototyping.
− Requires accurate governing equations (don’t have for turbulence,
combustion, etc..., which require modeling).
− Boundary conditions sometimes difficult to implement, e.g. outlet.
− Difficult to do in certain parameter regimes, e.g. high Reynolds numbers, and
complex geometries.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 4 / 454
Introduction to CFD

Experimental:
+ Easier to get overall quantities for problem (e.g. lift and drag on an airfoil).
+ No “modeling” necessary.
− Often requires intrusive measurement probes.
− Limited measurement accuracy.
− Limited measurement resolution.
− Effects of support apparatus, end walls, etc... must be considered.
− Some quantities difficult to obtain, e.g. streamfunction, vorticity, etc....
− Experimental equipment often expensive and takes up space.
− Difficult and costly to test full-scale models.
Note: Computational approaches do not replace analytical or experimental
approaches, but complement them.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 5 / 454

Numerical Methods: General Considerations and Approaches Components and Properties of a Numerical Solution

Outline

2 Numerical Methods: General Considerations and Approaches


Components and Properties of a Numerical Solution
Numerical Solution Approaches
Finite Difference
Boundary-Value Problems
Finite Element
Spectral Methods
Vortex Methods

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 6 / 454
Numerical Methods: General Considerations and Approaches Components and Properties of a Numerical Solution

Numerical Solution Procedure:

Physical System
i.e. Reality

1 Physical Laws + Models

Mathematical Model Analytical


i.e. Governing Equations Solution
(odes or pdes)

2 Discretization

System of Linear Equations

3 Matrix Solver

Numerical Solution

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 7 / 454

Numerical Methods: General Considerations and Approaches Components and Properties of a Numerical Solution

Steps:
1 Mass, momentum and energy conservation + models, idealizations, etc...
2 Discretization – approximation of the continuous differential equation(s) by a
system of algebraic equations for the dependent variables at discrete locations
in the independent variables (space and time). For example, using finite
differences

3 Numerical solutions of linear systems of equations.


⇒ Method of discretization often produces certain structure: e.g. second-order
centered differences lead to a tridiagonal system of equations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 8 / 454
Numerical Methods: General Considerations and Approaches Components and Properties of a Numerical Solution

Sources of errors in numerical solutions arise due to each step of the Numerical
Solution Procedure:
1 Modeling errors – difference between actual flow and exact solution of
mathematical model.
2 Discretization errors – difference between exact solution of governing
equations and exact solution of algebraic equations.
i) Method of discretization → inherent error of method, i.e. truncation error.
ii) Computational grid → can be refined.
3 Iterative convergence errors – difference between numerical solution, i.e.
iterative, and exact solution to algebraic equations.
This is where the “pragmatism” mentioned by Dawes is an important element of
CFD.
e.g. DNS of turbulence in other than elementary flows is currently impossible.
⇒ Compromises must often be made at each stage of the procedure.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 9 / 454

Numerical Methods: General Considerations and Approaches Components and Properties of a Numerical Solution

Properties of successful numerical solution methods:


1 Consistency – discretized equations must become governing equations as grid
size goes to zero
truncation error = discretized equations – exact equations
e.g.:
du ui+1 − ui−1
= + O(∆x2 ),
dx 2∆x
where O(∆x2 ) is the truncation error. Therefore, from the definition of the
−ui−1
derivative, as ∆x → 0, ui+12∆x → du
dx for consistency.
2 Stability – numerical procedure must not magnify errors produced in the
numerical solution.
Unstable method ⇒ diverge from exact solution.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 10 / 454
Numerical Methods: General Considerations and Approaches Components and Properties of a Numerical Solution

3 Convergence – numerical solution of discretized equations approaches exact


solution of governing equations as grid size goes to zero (cf. iterative
convergence).
Generally don’t have exact solution; therefore, refine grid until
grid-independent solution is obtained.
Truncation error gives the rate of convergence to the exact solution as grid is
reduced, e.g. for O(∆x2 ) ⇒ ∆x → e, where e is the error. Then halving the
grid size reduces the error by a factor of four.
4 “Correctness” – numerical solution should compare favorably with available
analytical solutions, experimental results or other computational solutions
within the limitations of each.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 11 / 454

Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Outline

2 Numerical Methods: General Considerations and Approaches


Components and Properties of a Numerical Solution
Numerical Solution Approaches
Finite Difference
Boundary-Value Problems
Finite Element
Spectral Methods
Vortex Methods

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 12 / 454
Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

The discretization step (step 2 in the Numerical Solution Procedure) may be


accomplished using different methods. The most common approaches used in
CFD are briefly described below.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 13 / 454

Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Finite Difference
Basic approach:
Discretize the governing equations in differential form using
Taylor-series-based finite-difference approximations at each grid point.
Produces algebraic equations involving each grid point and surrounding
points.
Local approximation method.
Popular in fluid dynamics research.
Advantages:
Relatively straightforward to understand and implement (based on Taylor
series).
Utilizes familiar differential form of governing equations.
Very general ⇒ Apply to a wide variety of problems (including complex
physics, e.g. fluids plus heat transfer plus combustion).
Can extend to higher-order approximations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 14 / 454
Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Disadvantages:
More difficult to implement for complex geometries.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 15 / 454

Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Finite Volume
Basic approach:
Apply conservation equations in integral form to a set of control volumes.
Produces algebraic equations for each control volume involving surrounding
control volumes.
Local approximation method.
Popular for commercial CFD codes (e.g. FLUENT).
Advantages:
Easier to treat complex geometries than finite-difference approach.
“Ensures” conservation of necessary quantities (i.e. mass, momentum,
energy, etc.), i.e. even if solution in inaccurate.
Disadvantages:
More difficult to construct higher-order schemes.
Uses less familiar integral formulation of governing equations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 16 / 454
Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Finite Element
Basic approach:
Apply conservation equations in variational form with weighting function to
set of finite elements.
Produces set of linear or nonlinear algebraic equations.
Local approximation method.
Popular in commercial codes (particularly for solid mechanics and heat
transfer).
Advantages:
Easy to treat complex geometries.
Disadvantages:
Results in unstructured grids.
Solution methods are inefficient for the types of matrices resulting from
finite-element discretizations (cf. finite difference ⇒ sparse, highly structured
matrices).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 17 / 454

Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Spectral Methods
Basic approach:
Solution of governing equations in differential form are approximated using
truncated (usually orthogonal) eigenfunction expansions.
Produces system of algebraic equations (steady) or system of ordinary
differential equations (unsteady) involving the coefficients in the
eigenfunction expansion.
Global approximation method.
Popular for direct numerical simulation (DNS) of turbulence.
Advantages:
Obtain highly accurate solutions when underlying solution is smooth.
Can achieve rapid convergence.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 18 / 454
Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Disadvantages:
Less straightforward to implement than finite difference.
More difficult to treat complicated boundary conditions (e.g. Neumann).
Small changes in problem can cause large changes in algorithm.
Not well suited for solutions having large gradients.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 19 / 454

Numerical Methods: General Considerations and Approaches Numerical Solution Approaches

Vortex Methods

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 20 / 454
Finite-Difference Methods Extended Fin Example

Outline

3 Finite-Difference Methods
Extended Fin Example
Formal Basis for Finite Differences
Application to Extended Fin Example
Properties of Tridiagonal Matrices
Thomas Algorithm
Extended Fin Example – Convection Boundary Condition

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 21 / 454

Finite-Difference Methods Extended Fin Example

As an example, consider the 1-D heat conduction in an extended surface with


convection to the ambient air.

The heat transfer within the extended fin is governed by the 1-D
ordinary-differential equation (see, for example, Incropera & DeWitt):

d2 T
   
1 dAc dT 1 h dAs
+ − (T − T∞ ) = 0, (3.1)
dx2 Ac dx dx Ac k dx

where

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 22 / 454
Finite-Difference Methods Extended Fin Example

T (x) = temperature distribution,


Ac (x) = cross-sectional area,
As (x) = surface area from base,
h = convective heat transfer coefficient,
k = thermal conductivity.
Note that equation (3.1) is a second-order ordinary differential equation with
variable coefficients, and it is a boundary-value problem requiring boundary
conditions at both ends of the domain.
Letting θ(x) = T (x) − T∞ , rewrite equation (3.1) as

d2 θ dθ
2
+ f (x) + g(x)θ = 0, (3.2)
dx dx
where
1 dAc
f (x) = ,
Ac dx

1 h dAs
g(x) = − .
Ac k dx

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 23 / 454

Finite-Difference Methods Extended Fin Example

For now, consider Dirichlet boundary conditions

θ = θb = Tb − T∞ at x = 0,
(3.3)
θ = θL = TL − T∞ at x = L.

Equation (3.2) with boundary conditions (3.3) represents the mathematical model
(step 1 in the Numerical Solution Procedure).
Step 2 → Discretization:
L
Divide the interval 0 ≤ x ≤ L into I equal subintervals of length ∆x = I:

Here, fi = f (xi ) and gi = g(xi ) are known, and the solution θi = θ(xi ) is to be
determined for i = 2, . . . , I.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 24 / 454
Finite-Difference Methods Extended Fin Example

In order to discretize the differential equation, consider the definition of the


derivative
dθ θ(xi + ∆x) − θ(xi )
= lim .
dx x=xi ∆x→0
∆x
This may be interpreted as a forward difference if ∆x is small, but not going to
zero, i.e. finite difference.
Graphical interpretation:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 25 / 454

Finite-Difference Methods Extended Fin Example


dθ θi+1 − θi
Forward Difference: ≈
dx xi
∆x

dθ θi − θi−1
Backward Difference: ≈
dx xi ∆x

dθ θi+1 − θi−1
Central Difference: ≈
dx xi
2∆x
Which approximation is better?

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 26 / 454
Finite-Difference Methods Formal Basis for Finite Differences

Outline

3 Finite-Difference Methods
Extended Fin Example
Formal Basis for Finite Differences
Application to Extended Fin Example
Properties of Tridiagonal Matrices
Thomas Algorithm
Extended Fin Example – Convection Boundary Condition

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 27 / 454

Finite-Difference Methods Formal Basis for Finite Differences

Consider the Taylor series expansion of a function θ(x) in the vicinity of the point
xi

(x − xi )2 d2 θ
   

θ(x) = θ(xi ) + (x − xi ) +
dx i 2! dx2 i
(3.4)
(x − xi )3 d3 θ (x − xi )n dn θ
   
+ + ··· + + ··· .
3! dx3 i n! dxn i

Apply the Taylor series at x = xi+1 : (xi+1 − xi = ∆x)

∆x2 d2 θ ∆x3 d3 θ ∆xn dn θ


       

θi+1 = θi +∆x + + +· · ·+ +· · · .
dx i 2 dx2 i 6 dx3 i n! dxn i
(3.5)
Solving for (dθ/dx)i gives

∆x d2 θ ∆xn−1 dn θ
     
dθ θi+1 − θi
= − − ··· − + ··· . (3.6)
dx i ∆x 2 dx2 i n! dxn i

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 28 / 454
Finite-Difference Methods Formal Basis for Finite Differences

Similarly, apply the Taylor series at x = xi−1 : (xi−1 − xi = −∆x)

∆x2 d2 θ ∆x3 d3 θ (−1)n ∆xn dn θ


       

θi−1 = θi −∆x + − +· · ·+ +· · · .
dx i 2 dx2 i 6 dx3 i n! dxn i
(3.7)
Solving again for (dθ/dx)i gives

θi − θi−1 ∆x d2 θ ∆x2 d3 θ (−1)n ∆xn−1 dn θ


       

= + − +· · ·+ +· · · .
dx i ∆x 2 dx2 i 6 dx3 i n! dxn i
(3.8)
Alternatively, subtract equation (3.7) from (3.5) to obtain

∆x3 d3 θ 2∆x2n+1 d2n+1 θ


     

θi+1 − θi−1 = 2∆x + +···+ +··· ,
dx i 3 dx3 i (2n + 1)! dx2n+1 i
(3.9)
and solve for (dθ/dx)i to obtain

θi+1 − θi−1 ∆x2 d3 θ ∆x2n


     2n+1 
dθ d θ
= − 3
−· · ·− 2n+1
−· · · . (3.10)
dx i 2∆x 6 dx i (2n + 1)! dx i

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 29 / 454

Finite-Difference Methods Formal Basis for Finite Differences

Equations (3.6), (3.8) and (3.10) are exact expressions for the first derivative
(dθ/dx)i , i.e. if all of the terms are retained in the expansions.
Approximate finite difference expressions for the first derivative may then be
obtained by truncating the series after the first term:
 
dθ θi+1 − θi
≈ + O(∆x) → Forward difference
dx i ∆x
 
dθ θi − θi−1
≈ + O(∆x) → Backward difference
dx i ∆x
 
dθ θi+1 − θi−1
≈ + O(∆x2 ) → Central difference
dx i 2∆x

The truncation error is the sum of all the truncated terms.


⇒ For small ∆x, successive terms get smaller and the order of the truncation
error is given by the first truncated term.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 30 / 454
Finite-Difference Methods Formal Basis for Finite Differences

Higher-order approximations may be obtained by various manipulations of the


Taylor series at additional points. For example, to obtain a second-order acccurate
forward-difference approximation to the first derivative, apply the Taylor series at
xi+2 as follows

(2∆x)2 d2 θ (2∆x)3 d3 θ
     

θi+2 = θi + 2∆x + + + · · · . (3.11)
dx i 2! dx2 i 3! dx3 i

We eliminate the (d2 θ/dx2 )i term by taking 4×(3.5) − (3.11) to obtain

2∆x3 d3 θ
   

4θi+1 − θi+2 = 3θi + 2∆x − + ··· .
dx i 3 dx3 i

Solving for (dθ/dx)i gives

∆x2 d3 θ
   
dθ −3θi + 4θi+1 − θi+2
= + + ··· , (3.12)
dx i 2∆x 3 dx3 i

which is second-order accurate and involves the point of interest and the two
previous points.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 31 / 454

Finite-Difference Methods Formal Basis for Finite Differences

For a second-order accurate central-difference approximation to the second


derivative add equation (3.5) and (3.7) for θi+1 and θi−1 , respectively, to
eliminate the (dθ/dx)i term. This gives
 2 
∆x4 d4 θ
 
2 d θ
θi+1 + θi−1 = 2θi + ∆x + + ··· .
dx2 i 12 dx4 i

Solving for (d2 θ/dx2 )i leads to


 2 
∆x2 d4 θ
 
d θ θi+1 − 2θi + θi−1
= − + ··· . (3.13)
dx2 i ∆x2 12 dx4 i

Note that the second-order accurate central-difference approximations for both


the first and second derivatives involve the point of interest and its two neighbors.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 32 / 454
Finite-Difference Methods Application to Extended Fin Example

Outline

3 Finite-Difference Methods
Extended Fin Example
Formal Basis for Finite Differences
Application to Extended Fin Example
Properties of Tridiagonal Matrices
Thomas Algorithm
Extended Fin Example – Convection Boundary Condition

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 33 / 454

Finite-Difference Methods Application to Extended Fin Example

Returning to the fin equation (3.2)


d2 θ dθ
+ f (x) + g(x)θ = 0,
dx2 dx
and approximating the derivatives using second-order accurate finite differences
gives
θi+1 − 2θi + θi−1 θi+1 − θi−1
+ fi + gi θi = 0,
(∆x)2 2∆x
for the point xi . Multiplying by (∆x)2 and collecting terms leads to
ai θi−1 + bi θi + ci θi+1 = di , i = 2, . . . , I, (3.14)
where the difference equation is applied at each interior point of the domain, and
∆x
ai = 1 − fi ,
2

bi = −2 + (∆x)2 gi ,

∆x
ci = 1+ fi ,
2

di = 0.
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 34 / 454
Finite-Difference Methods Application to Extended Fin Example

Note that because we have discretized the differential equation at each interior
grid point, we obtain a set of (I − 1) algebraic equations for the (I − 1) unknown
values of the temperature θi , i = 2, . . . , I at each interior grid point.
The coefficient matrix for the difference equation (3.14) is tridiagonal
    
b2 c2 0 0 ··· 0 0 0 θ2 d2 − a2 θ1
a3 b3
 c3 0 · · · 0 0 0    θ3  
   d3 

 0 a4 b4 c4 · · · 0 0 0   θ4   d 4

..   ..  = 
    
 .. .. .. .. . . .. .. .. 
.
 . . . . . . . 

 .  
  . 

0 0 0 0 · · · aI−1 bI−1 cI−1  θI−1   dI−1 
0 0 0 0 ··· 0 aI bI θI dI − cI θI+1

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 35 / 454

Finite-Difference Methods Properties of Tridiagonal Matrices

Outline

3 Finite-Difference Methods
Extended Fin Example
Formal Basis for Finite Differences
Application to Extended Fin Example
Properties of Tridiagonal Matrices
Thomas Algorithm
Extended Fin Example – Convection Boundary Condition

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 36 / 454
Finite-Difference Methods Properties of Tridiagonal Matrices

Consider a N × N tridiagonal matrix with constant ai , bi , and ci


 
b c 0 ··· 0 0
a b c
 · · · 0 0 
0 a b · · · 0 0
A = . . . .. ..  .
 
 .. .. .. ..
 . . . 
0 0 0 · · · b c
0 0 0 ··· a b

It can be shown that the eigenvalues of a tridiagonal matrix are



 

λj = b + 2 ac cos , j = 1, . . . , N. (3.15)
N +1
The eigenvalues with the largest and smallest magnitudes are (which is largest or
smallest depends upon a, b and c)

 
π
|λ1 | = b + 2 ac cos ,
N +1

 
N π
|λN | = b + 2 ac cos .
N +1
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 37 / 454

Finite-Difference Methods Properties of Tridiagonal Matrices

Then the condition number of A using an L2 norm is defined by

|λ|max
cond2 (A) = .
|λ|min

Let us consider N large. Thus, expanding cosine in a Taylor series (the first is
expanded about π/(N + 1) → 0 and the second about N π/(N + 1) → π
  2  4

π π 1 1 π
cos =1− + + ··· ,
N +1 N +1 2!4! N + 1
   2  2
Nπ 1 Nπ 1 π
cos = −1 + − π − · · · = −1 + − ··· .
N +1 2! N + 1 2 N +1

Consider the common case that may result from the use of central differences for
a second-order derivative:

a = 1, b = −2, c = 1,

which is weakly diagonally dominant.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 38 / 454
Finite-Difference Methods Properties of Tridiagonal Matrices

Then with the large N expansions from above


"  2 #  2
p 1 π π
|λ1 | = −2 + 2 (1)(1) 1 − + ··· = + ··· ,

2 N +1 N +1
"  2 #
p 1 π
|λN | = −2 + 2 (1)(1) −1 + − ··· = 4 + ··· .

2 N +1
(3.16)
Thus, the condition number for large N is approximately

4 4(N + 1)2
cond2 (A) ≈ = , for large N .
(π/(N + 1))2 π2

⇒ Increases proportional to N 2 with increasing N .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 39 / 454

Finite-Difference Methods Properties of Tridiagonal Matrices

Now consider the case with

a = 1, b = −4, c = 1,

which is strictly, or strongly, diagonally dominant. Then from equation (3.16), the
condition number for large N is approximately
6
cond2 (A) ≈ = 3, for large N .
2

⇒ Constant with increasing N .


It can be
Pproven that for any strictly diagonally dominant matrix A, i.e.
N
|aii | > j=1,j6=i |aij |:
1) A is nonsingular; therefore, it is invertible.
2) Gauss elimination does not require row interchanges for conditioning.
3) Computations are stable with respect to round-off errors.
To see the influence of condition number on round-off error, see the Mathematica
notebook “Ill-Conditioned Matrices and Round-Off Error.”

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 40 / 454
Finite-Difference Methods Thomas Algorithm

Outline

3 Finite-Difference Methods
Extended Fin Example
Formal Basis for Finite Differences
Application to Extended Fin Example
Properties of Tridiagonal Matrices
Thomas Algorithm
Extended Fin Example – Convection Boundary Condition

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 41 / 454

Finite-Difference Methods Thomas Algorithm

Tridiagonal systems of equations may be solved directly using the Thomas


algorithm, which is essentially application of Gauss elimination and consists of two
stages (shown here for Dirichlet boundary conditions):
1) Forward Elimination:

F1 = 0, δ1 = θ1 = θb = boundary condition
ci di − ai δi−1
Fi = , δi = , i = 2, . . . , I.
bi − ai Fi−1 bi − ai Fi−1

2) Back Substitution:

θI+1 = θL = boundary condition


θi = δi − Fi θi+1 , i = I, . . . , 2.

Note the order of evaluation for the back substitution.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 42 / 454
Finite-Difference Methods Thomas Algorithm

Notes:
1 See Anderson, Appendix A for a derivation.
2 The Thomas algorithm only requires O(I) operations, which is as good a
scaling as one could hope for. Gauss elimination of a full (dense) matrix, for
example, requires O(I 3 ) operations.
3 To prevent ill-conditioning the system of equations should be diagonally
dominant. For our tridiagonal system, this means

|bi | ≥ |ai | + |ci |,

where if the greater than sign applies we say that the matrix is strictly
diagonally dominant, or weakly diagonally dominant if the equal sign applies.
Performing operations with ill-conditioned matrices can result in the growth
of small round-off errors that then contaminate the solution. For example,
note how errors could accumulate in the Fi , δi coefficients in the Thomas
algorithm.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 43 / 454

Finite-Difference Methods Extended Fin Example – Convection Boundary Condition

Outline

3 Finite-Difference Methods
Extended Fin Example
Formal Basis for Finite Differences
Application to Extended Fin Example
Properties of Tridiagonal Matrices
Thomas Algorithm
Extended Fin Example – Convection Boundary Condition

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 44 / 454
Finite-Difference Methods Extended Fin Example – Convection Boundary Condition

Rather than a Dirichlet boundary condition at the tip of the fin, which assumes
that we know the temperature there, let us consider a more realistic convection
condition at the tip
dT
−k = h(T − T∞ ) at x = L,
dx
or with θ(x) = T (x) − T∞
 

−k = hθ(L)
dx x=L

Note that the convection condition results in specifying a linear combination of


the temperature and its derivative at the tip. This is known as a Robin boundary
condition and is of the form

pθ + q =r at x = L,
dx
where p = h, q = k, r = 0, which are inputs to the Thomas algorithm.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 45 / 454

Finite-Difference Methods Extended Fin Example – Convection Boundary Condition

Now let us consider evaluation of the heat flux at the base of the fin
   
dT dθ
qb = −kAc (0) = −kAc (0) .
dx x=0 dx x=0

In order to evaluate dθ/dx at the base x = 0, we must use a forward difference.


From equation (3.6) applied at i = 1, we have the first-order accurate
forward-difference approximation
 
dθ θ2 − θ1
≈ + O(∆x).
dx x=0 ∆x

For a more accurate approximation, we may use the second-order accurate


approximation from equation (3.12) applied at i = 1
 
dθ −3θ1 + 4θ2 − θ3
≈ + O(∆x2 ).
dx x=0 2∆x

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 46 / 454
Finite-Difference Methods Extended Fin Example – Convection Boundary Condition

Even higher-order finite-difference approximations may be formed. For example,


the third-order, forward-difference approximation is
 
dθ −11θ1 + 18θ2 − 9θ3 + 2θ4
≈ + O(∆x3 ).
dx x=0 6∆x

Observe that each successive approximation requires one additional point in the
interior of the domain.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 47 / 454

Classification of Second-Order Partial Differential Equations Mathematical Classification

Outline

4 Classification of Second-Order Partial Differential Equations


Mathematical Classification
Hyperbolic Equations
Parabolic Equations
Elliptic Equations
Mixed Equations

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 48 / 454
Classification of Second-Order Partial Differential Equations Mathematical Classification

Consider the general second-order partial differential equation for u(x, y)

auxx + buxy + cuyy + dux + euy + f u = g. (4.1)

Linear ⇒ a, b, c, d, e and f are only functions of (x, y).


Quasi-linear ⇒ a, b and c may be functions of (x, y, u, ux , uy ), i.e. equation is
linear in highest derivative (e.g. no u2xx terms).
Let us determine the criteria necessary for the existence of a smooth
(differentiable) and unique (single-valued) solution along a curve C:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 49 / 454

Classification of Second-Order Partial Differential Equations Mathematical Classification

Along C, let
φ1 (τ ) = uxx , φ2 (τ ) = uxy , φ3 (τ ) = uyy ,
(4.2)
ψ1 (τ ) = ux , ψ2 (τ ) = uy .
Substituting into equation (4.1) gives

aφ1 + bφ2 + cφ3 = g − dψ1 − eψ2 − f u = H. (4.3)

Transforming from (x, y) → τ , we have

d dx ∂ dy ∂
= + .
dτ dτ ∂x dτ ∂y

Thus, from (4.2)

dψ1 d dx dy dx dy
= ux = uxx + uxy = φ1 + φ2 , (4.4)
dτ dτ dτ dτ dτ dτ
dψ2 d dx dy dx dy
= uy = uxy + uyy = φ2 + φ3 . (4.5)
dτ dτ dτ dτ dτ dτ

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 50 / 454
Classification of Second-Order Partial Differential Equations Mathematical Classification

Equations (4.3)–(4.5) are three equations for three unknowns, i.e. the
second-order derivatives φ1 , φ2 and φ3 . Written in matrix form, they are
    
a b c φ1 H
 dx/dτ dy/dτ 0   φ2  =  dψ1 /dτ 
0 dx/dτ dy/dτ φ3 dψ2 /dτ

If the determinant of the coefficient matrix is not equal to zero, a unique solution
exists for the second derivatives along the curve C. It can be shown that if the
second-order derivatives exist, then derivatives of all orders exist along C as well.
On the other hand, if the determinant of the coefficient matrix is equal to zero,
the solution is not unique, i.e. the second derivatives are discontinuous along C.
Setting the determinant equal to zero gives
 2  2   
dy dx dx dy
a +c −b = 0,
dτ dτ dτ dτ

or multiplying by (dτ /dx)2


 2  
dy dy
a −b + c = 0.
dx dx
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 51 / 454

Classification of Second-Order Partial Differential Equations Mathematical Classification

This is a quadratic equation for dy/dx, which is the slope of the curve C.
Thus, √
dy b ± b2 − 4ac
= . (4.6)
dx 2a
→ The curves C for which y(x) satisfy (4.6) are called characteristic curves of
equation (4.1), and they are curves along which the second-order derivatives
are discontinuous.
Because the characteristics must be real, their behavior is determined by the sign
of b2 − 4ac:
b2 − 4ac > 0 ⇒ 2 real roots ⇒ 2 characteristics ⇒ hyperbolic p.d.e.
b2 − 4ac = 0 ⇒ 1 real root ⇒ 1 characteristic ⇒ parabolic p.d.e.
b2 − 4ac < 0 ⇒ no real roots ⇒ no characteristics ⇒ elliptic p.d.e.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 52 / 454
Classification of Second-Order Partial Differential Equations Mathematical Classification

The terminology arises from classification of the second-degree algebraic equation


ax2 + bxy + cy 2 + dx + ey + f = 0, i.e. conic sections.
Note that if from equation (4.1), we write (let x = x1 , y = x2 , then
au11 + 2b u12 + 2b u21 + cu22 = H)
 
a b/2
A=
b/2 c

(analogous to quadratic forms in MMAE 501). Then

det[A] = ac − b2 /4

or 
 > 0, hyperbolic
2
−4 det[A] = b − 4ac = 0, parabolic
< 0, elliptic

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 53 / 454

Classification of Second-Order Partial Differential Equations Mathematical Classification

Notes:
1 Physically, characteristics are curves along which information propagates in
the solution.
2 For the case of elliptic equations, the matrix A is positive definite.
3 The classification depends on the coefficients of the highest-order derivatives,
i.e. a, b and c.
4 It can be shown that the classification of a partial differential equation is
independent of the coordinate system (see, for example, Tannehill et al.).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 54 / 454
Classification of Second-Order Partial Differential Equations Hyperbolic Equations

Outline

4 Classification of Second-Order Partial Differential Equations


Mathematical Classification
Hyperbolic Equations
Parabolic Equations
Elliptic Equations
Mixed Equations

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 55 / 454

Classification of Second-Order Partial Differential Equations Hyperbolic Equations

From equation (4.6), for a hyperbolic equation there are two real roots, say

dy dy
= λ1 , = λ2 . (4.7)
dx dx
If a, b and c are constant (λ1 , λ2 constant), then we may integrate to obtain

y = λ1 x + x1 , y = λ 2 x + x2 ,

which are straight lines. Therefore, the solution propagates along two linear
characteristic curves.
For example, consider the wave equation

∂2u 2
2∂ u
= σ , (y → t : a = σ 2 , b = 0, c = −1)
∂t2 ∂x2
where u(x, t) is the amplitude of the wave, and σ is the wave speed. Therefore,
from equations (4.6) and (4.7),
1 1
λ1 = , λ2 = − .
σ σ
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 56 / 454
Classification of Second-Order Partial Differential Equations Hyperbolic Equations

Therefore, the characteristics of the wave equation with a, b, and c constant are
straight lines with slopes 1/σ and −1/σ.

The solution to the wave equation is of the form

u(x, t) = F1 (x + σt) + F2 (x − σt),

and initial conditions are required at say t = 0, such as

u(x, 0) = f (x), ut (x, 0) = g(x).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 57 / 454

Classification of Second-Order Partial Differential Equations Hyperbolic Equations

Note that no boundary conditions are necessary at specified values of x, i.e. the
solution (F1 , F2 ) only depends upon the initial conditions.
Hyperbolic equations in fluid dynamics:
Unsteady inviscid flow
Steady supersonic inviscid flow

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 58 / 454
Classification of Second-Order Partial Differential Equations Parabolic Equations

Outline

4 Classification of Second-Order Partial Differential Equations


Mathematical Classification
Hyperbolic Equations
Parabolic Equations
Elliptic Equations
Mixed Equations

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 59 / 454

Classification of Second-Order Partial Differential Equations Parabolic Equations

From equation (4.6), for a parabolic equation there is only one real root, which is

dy b
= . (4.8)
dx 2a
If a and b are constant, then we may integrate to obtain
b
y= x + γ1 ,
2a
which is a straight line. Therefore, the solution propagates along one linear
characteristic direction (usually time).
For example, consider the one-dimensional, unsteady diffusion equation (e.g. heat
conduction)
∂u ∂2u
= α 2 , (y → t : a = α, b = c = 0)
∂t ∂x
where u(x, t) is the quantity undergoing diffusion (e.g. temperature), and α is the
diffusivity.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 60 / 454
Classification of Second-Order Partial Differential Equations Parabolic Equations

The solution marches forward in time, i.e. the characteristics (with b = 0) are lines
of constant t.

Initial and boundary conditions are required, such as

u(x, 0) = u0 (x), u(x1 , t) = f (t), u(x2 , t) = g(t).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 61 / 454

Classification of Second-Order Partial Differential Equations Parabolic Equations

Parabolic equations in fluid dynamics:


Unsteady Navier-Stokes
Steady boundary-layer flow (no separation):

∂u ∂u ∂p ∂ 2 u
u +v =− + , (a = 0, b = 0, c = 1).
∂x ∂y ∂x ∂y 2
In this case the solution marches forward in the x−direction from an initial
velocity profile.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 62 / 454
Classification of Second-Order Partial Differential Equations Elliptic Equations

Outline

4 Classification of Second-Order Partial Differential Equations


Mathematical Classification
Hyperbolic Equations
Parabolic Equations
Elliptic Equations
Mixed Equations

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 63 / 454

Classification of Second-Order Partial Differential Equations Elliptic Equations

Disturbances have infinite speed of propagation in all directions, i.e. no


characteristics ⇒ A disturbance anywhere affects the solution everywhere
instantaneously.
For example, consider the Laplace equation (e.g. two-dimensional steady heat
conduction, potential flow, etc.)

∂2u ∂2u
+ 2 = 0, (a = 1, b = 0, c = 1)
∂x2 ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 64 / 454
Classification of Second-Order Partial Differential Equations Elliptic Equations

Requires a global solution strategy, and boundary conditions (typically u or


∂u/∂n) must be specified on a closed contour bounding the domain.
Elliptic equations in fluid dynamics:
Steady Navier-Stokes (also steady energy equation)
Potential (inviscid, incompressible, irrotational) flow
Poisson equation for pressure or streamfunction

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 65 / 454

Classification of Second-Order Partial Differential Equations Mixed Equations

Outline

4 Classification of Second-Order Partial Differential Equations


Mathematical Classification
Hyperbolic Equations
Parabolic Equations
Elliptic Equations
Mixed Equations

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 66 / 454
Classification of Second-Order Partial Differential Equations Mixed Equations

If a, b and c are variable coefficients, then b2 − 4ac may change sign with space
and/or time.
⇒ Character of equations may be different in certain regions.
For example, consider transonic flow (Mach ∼ 1). The governing equation for
two-dimensional, steady, compressible, potential flow about a slender body is

∂2φ ∂2φ
(1 − M 2 ) + = 0, (x → s, y → n : a = 1 − M 2 , b = 0, c = 1),
∂s2 ∂n2
where φ(s, n) is the velocity potential, M is the local Mach number, and s and n
are streamline coordinates with s locally being tangent to the streamline and n
being normal to the streamline.
To determine the nature of the equation, observe that

b2 − 4ac = 0 − 4(1 − M 2 )(1) = −4(1 − M 2 );

therefore,
M < 1 ⇒ b2 − 4ac < 0 ⇒ Elliptic
M = 1 ⇒ b2 − 4ac = 0 ⇒ Parabolic
M > 1 ⇒ b2 − 4ac > 0 ⇒ Hyperbolic
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 67 / 454

Classification of Second-Order Partial Differential Equations Mixed Equations

In the above example, we have the same equation, but different behavior in
various regions. In the following example, we have different equations in different
regions of the flow.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 68 / 454
Classification of Second-Order Partial Differential Equations Mixed Equations

Consider the steady, incompressible viscous flow past a surface.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 69 / 454

Numerical Solutions of Elliptic Problems Introduction

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 70 / 454
Numerical Solutions of Elliptic Problems Introduction

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 71 / 454

Numerical Solutions of Elliptic Problems Introduction

Recall that elliptic problems have no preferred direction of propagation; therefore,


they require:
1 Boundary conditions on a closed contour surrounding the domain, i.e.
boundary-value problem.

2 Global solution method.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 72 / 454
Numerical Solutions of Elliptic Problems Introduction

Types of boundary conditions:


1 Dirichlet – φ specified.
∂φ
2 Neumann – ∂n specified (n is normal to boundary).
∂φ
3 Robin (mixed) – ∂n + aφ = b specified
Notes:
1 Combinations of the above boundary conditions may be applied on different
portions of the boundary as long as some boundary condition is applied at
every point along the boundary contour.
2 Solutions to the Laplace and Poisson equations with Neumann boundary
conditions on the entire boundary can only be determined relative to an
unknown constant, i.e. φ(x, y) + c is a solution.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 73 / 454

Numerical Solutions of Elliptic Problems Introduction

Linear vs. Non-Linear Elliptic Problems:


Linear – Linear equation and boundary conditions
e.g. Laplace (f = 0) or Poisson equations

∂2φ ∂2φ
+ = f (x, y),
∂x2 ∂y 2
with Dirichlet, Neumann or Robin boundary conditions.
Non-Linear:
1 Linear equation with non-linear boundary conditions; e.g. heat conduction with
radiation condition
∂2T ∂2T
+ = 0,
∂x2 ∂y 2
∂T
= D(T 4 − Tsur4
).
∂n
2 Non-linear equation; e.g.: Navier-Stokes equations
„ 2
∂2u
«
∂u ∂u ∂p 1 ∂ u
u +v =− + + .
∂x ∂y ∂x Re ∂x2 ∂y 2

Note that as Reynolds number, Re, increases, the non-linearity increases.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 74 / 454
Numerical Solutions of Elliptic Problems Finite-Difference Methods for the Poisson Equation

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 75 / 454

Numerical Solutions of Elliptic Problems Finite-Difference Methods for the Poisson Equation

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 76 / 454
Numerical Solutions of Elliptic Problems Finite-Difference Methods for the Poisson Equation

As a representative elliptic equation, which is important in its own right, consider


the two-dimensional Poisson equation

∂2φ ∂2φ
+ 2 = f (x, y), (5.1)
∂x2 ∂y

where if f (x, y) = 0 we have the Laplace equation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 77 / 454

Numerical Solutions of Elliptic Problems Finite-Difference Methods for the Poisson Equation

In order to define the grid:


Divide the x interval, 0 ≤ x ≤ a, into I equal subintervals of length ∆x, with
xi = ∆x(i − 1).
Divide the y interval, 0 ≤ y ≤ b, into J equal subintervals of length ∆y, with
yj = ∆y(j − 1).

⇒ 2-D grid intersecting in (I + 1) × (J + 1) points.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 78 / 454
Numerical Solutions of Elliptic Problems Finite-Difference Methods for the Poisson Equation

Consider an approximation to equation (5.1) at a typical point (i, j); the five
point finite-difference stencil is

Using second-order accurate, central differences (with φi,j = φ(xi , yj )):

∂2φ φi+1,j − 2φi,j + φi−1,j


2
≈ + O(∆x2 ),
∂x (∆x)2

∂2φ φi,j+1 − 2φi,j + φi,j−1


2
≈ + O(∆y 2 ).
∂y (∆y)2

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 79 / 454

Numerical Solutions of Elliptic Problems Finite-Difference Methods for the Poisson Equation

Substituting into (5.1) and multiplying by (∆x)2 gives the final form of the
finite-difference equation
"  2 #  2
∆x ∆x
φi+1,j − 2 1 + φi,j + φi−1,j + (φi,j+1 + φi,j−1 ) = (∆x)2 fi,j ,
∆y ∆y
(5.2)
which results in a system of (I + 1) × (J + 1) equations for (I + 1) × (J + 1)
unknowns.
There are two options for solving such systems of equations:
1 Direct Methods:
i) No iterative convergence errors.
ii) Efficient for certain types of linear systems, e.g. tridiagonal, block-tridiagonal.
iii) Become less efficient for large systems of equations.
iv) Typically cannot adapt to non-linear problems.
2 Iterative Methods:
i) Iterative convergence errors.
ii) Generally more efficient for large systems of equations.
iii) Apply to non-linear problems.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 80 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 81 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 82 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

We wish to consider direct methods for solving the discretized Poisson equation.
Repeating the difference equation (5.2) for the Poisson equation, we have
¯ i,j + φi−1,j + ∆
φi+1,j − 2(1 + ∆)φ ¯ (φi,j+1 + φi,j−1 ) = (∆x)2 fi,j , (5.3)
¯ = (∆x/∆y)2 , and i = 1, . . . , I + 1; j = 1, . . . , J + 1.
where ∆
In order to write the system of difference equations in matrix form, let us renumber
the two-dimensional mesh (i, j) into a one-dimensional array (n) as follows

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 83 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Thus, the relationship between the one-dimensional and two-dimensional indices is

n = (i − 1)(J + 1) + j, n = 1, . . . , (I + 1)(J + 1),

where i = 1, . . . , I + 1 and j = 1, . . . , J + 1.
Therefore, our five-point finite difference stencil becomes

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 84 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

And the finite-difference equation (5.4) for the Poisson equation becomes (with
∆ = ∆x = ∆y ⇒ ∆ ¯ = 1)

φn+J+1 + φn−(J+1) + φn+1 + φn−1 − 4φn = ∆2 fn , (5.4)

where n = 1, . . . , (I + 1)(J + 1).


Thus, the system of equations Aφ = d is of the form
    
D I 0 ··· 0 0 φ1 d1
 I D I ··· 0 0 φ2   d2 
    
0 I D ··· 0 0 φ 3
  d 3

= .
    
 .. .. .. . . .. ..   ..   ..
.
 . . . . . 
  . 
 
 . 

 0 0 0 · · · D I  φ(I+1)(J+1)−1  d(I+1)(J+1)−1 
0 0 0 ··· I D φ(I+1)(J+1) d(I+1)(J+1)

A is a block tridiagonal matrix with (I + 1) blocks in both directions, where each


block is a (J + 1) × (J + 1) matrix, and d contains all known information from
the boundary conditions.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 85 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

0 is a zero matrix block, I is an identity matrix block, and the tridiagonal blocks
are  
−4 1 0 ··· 0 0
 1 −4 1 · · · 0 0
 
0 1 −4 · · · 0 0 
D= . .
 
. . . . .
 .. .. .. .. .. .. 

 
0 0 0 · · · −4 1 
0 0 0 ··· 1 −4
In general, the method for solving the system Aφ = d depends upon the form of
the coefficient matrix A:
Full or dense ⇒ Gauss elimination, LU decomposition, etc. (very expensive
computationally).
Sparse and banded → Result from discretizations of certain classes of
problems (e.g. separable elliptic partial differential equations, such as the
Poisson equation):
→ Tridiagonal ⇒ Thomas algorithm
→ Fast Fourier Transform (FFT) and/or cyclic reduction, which generally are the
fastest methods for problems in which they apply.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 86 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Fourier Transform Methods


The Fourier transform of a continuous function h(t) is defined by
Z ∞
H(f ) = h(t)e2πif t dt,
−∞

where f is the frequency, and i is the imaginary number. The inverse transform is
Z ∞
h(t) = H(f )e−2πif t df.
−∞

Consider the discrete form of the Fourier transform in which we have N values of
h(t) at discrete points defined by

hk = h(tk ), tk = k∆, k = 0, 1, 2, . . . , N − 1,

where ∆ is the step size in the data.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 87 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

After taking the Fourier transform, we will have N discrete points in the
frequency domain defined by
n N N
fn = , n=− ,..., ,
N∆ 2 2
corresponding to the Nyquist critical frequency range. Thus, the discrete Fourier
transform is Z ∞
H(fn ) = h(t)e2πifn t dt,
−∞

or approximating the integral by summing over each ∆ interval


N
X −1
H(fn ) ≈ hk e2πifn tk ∆.
k=0

Using the fact that fn tk = (n/N ∆)k∆ = nk/N gives


N
X −1
H(fn ) ≈ ∆ hk e2πikn/N .
k=0

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 88 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Denoting the summation in the previous expression by ĥn , we write


N
X −1
∴ H(fn ) ≈ ∆ĥn , ĥn = hk e2πikn/N . (5.5)
k=0

The discrete inverse Fourier transform is then


N/2
1 X
hk = ĥn e−2πikn/N . (5.6)
N
n=−N/2

The above expressions are all for a one-dimensional domain.


For a two-dimensional grid, we define:
k = 0, . . . , K − 1, l = 0, . . . , L − 1 → Physical grid
m = −K K
2 ,..., 2 , n = − L2 , . . . , L2 → Transform space (frequency)

The two-dimensional discrete Fourier transform of φ is (analogous to (5.5)


K−1
X L−1
X
2
Φ(fm,n ) ≈ ∆ φ̂m,n , φ̂m,n = φk,l e2πimk/K e2πinl/L . (5.7)
k=0 l=0
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 89 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

This is equivalent to taking the Fourier transform in each direction. The inverse
Fourier transform is (analogous to (5.6))
K/2 L/2
1 X X
φk,l = φ̂m,n e−2πikm/K e−2πiln/L . (5.8)
KL
m=−K/2 n=−L/2

Now let us apply this to the discretized Poisson equation (5.2). For simplicity, set
∆ = ∆x = ∆y giving (with i → k, j → l)

φk+1,l + φk−1,l + φk,l+1 + φk,l−1 − 4φk,l = ∆2 fk,l , (5.9)

where now fk,l is the right-hand-side of the Poisson equation (not frequency).
Substituting equation (5.8) into equation (5.9) leads to

φ̂m,n e−2πi(k+1)m/K e−2πiln/L + e−2πi(k−1)m/K e−2πiln/L




+e−2πikm/K e−2πi(l+1)n/L + e−2πikm/K e−2πi(l−1)n/L


−4e−2πikm/K e−2πiln/L = ∆2 fˆm,n e−2πikm/K e−2πiln/L ,


where fˆm,n is the Fourier transform of the right-hand-side fk,l .


MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 90 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Canceling the common factor e−2πikm/K e−2πiln/L , we have


h i
φ̂m,n e −2πim/K
+e 2πim/K
+e −2πin/L
+e 2πin/L
− 4 = ∆2 fˆm,n .

Recalling that cos(ax) = eiax + e−iax /2 gives




     
2πm 2πn
φ̂m,n 2 cos + 2 cos − 4 = ∆2 fˆm,n .
K L

Solving for φ̂m,n leads to

∆2 fˆm,n
φ̂m,n =  , (5.10)
2 cos 2πm 2πn
 
K + cos L −2

for m = 1, . . . , K − 1; n = 1, . . . , L − 1.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 91 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Therefore, to solve the difference equation (5.9) using Fourier transform methods:
1) Compute the Fourier transform fˆm,n of the right-hand-side fk,l using (similar
to (5.7))
K−1
X L−1
X
ˆ
fm,n = fk,l e2πimk/K e2πinl/L . (5.11)
k=0 l=0

2) Compute φ̂m,n from equation (5.10).


3) Compute φk,l using the inverse Fourier transform (5.8).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 92 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Notes:
1) The above procedure works for periodic boundary conditions, i.e. the solution
satisfies
φk,l = φk+K,l = φk,l+L .
For Dirichlet boundary conditions ⇒ Use sine transform.
For Neumann boundary conditions ⇒ Use cosine transform.
2) In practice, the Fourier (and inverse) transforms are computed using a Fast
Fourier Transform (FFT) technique (see, for example, Numerical Recipes).
3) Fourier transform methods can only be applied to partial differential
equations with constant coefficients in the direction(s) for which the Fourier
transform is applied.
4) We use Fourier transforms to solve the difference equation, not the
differential equation; therefore, this is not a spectral method.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 93 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Cyclic Reduction
Again, consider the Poisson equation

∂2φ ∂2φ
+ 2 = f (x, y),
∂x2 ∂y
discretized on a two-dimensional grid with ∆ = ∆x = ∆y and
i = 0, . . . , I; j = 0, . . . , J, where I = 2n with integer n ((I + 1) × (J + 1) points):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 94 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Applying central differences to the Poisson equation, the difference equation for
constant x-lines becomes

ui−1 − 2ui + ui+1 + B0 ui = ∆2 fi , i = 0, . . . , I, (5.12)

where
     
φi,0 fi,0 −2 1 0 ··· 0 0



 φi,1 


fi,1 1
 −2 1 ··· 0 0

  φi,2  fi,2 0 1 −2 ··· 0 0
0
ui =  , fi =  , B = . .
     
.. .. .. .. .. .. ..
  .   .  .. . . . . .

     
φi,J−1  fi,J−1  0 0 0 ··· −2 1
φi,J fi,J 0 0 0 ··· 1 −2

The first three terms in equation (5.12) correspond to the central difference in the
x-direction, and the fourth term corresponds to the central difference in the
y-direction (see B0 ).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 95 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Taking B = −2I + B0 , where I is the identity matrix, equation (5.12) becomes

ui−1 + Bui + ui+1 = ∆2 fi , i = 0, . . . , I, (5.13)

where the (J + 1) × (J + 1) matrix B is


 
−4 1 0 ··· 0 0
 1 −4 1 ··· 0 0
 
0 1 −4 ··· 0 0
B= . .
 
.. .. .. .. ..
 .. . . . . .

 
0 0 0 ··· −4 1
0 0 0 ··· 1 −4

Note that equation (5.13) corresponds to the block-tridiagonal matrix for equation
(5.4), where B is the tridiagonal portion and ui−1 and uu+1 are the ’fringes.’
Writing three successive equations of (5.13) for i − 1, i and i + 1:

ui−2 + Bui−1 + ui = ∆2 fi−1 ,


ui−1 + Bui + ui+1 = ∆2 fi ,
ui + Bui+1 + ui+2 = ∆2 fi+1 .
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 96 / 454
Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

Multiplying −B times the middle equation and adding all three gives

ui−2 + B∗ ui + ui+2 = ∆2 fi∗ , (5.14)

where
B∗ = 2I − B2 ,
fi∗ = fi−1 − Bfi + fi+1 .
This is an equation of the same form as (5.13); therefore, applying this procedure
to all even numbered i equations in (5.13) reduces the number of equations by a
factor of two.
This cyclic reduction procedure can be repeated until a single equation remains for
the middle line of variables, uI/2 (I = 2n , with integer n), which is tridiagonal.
Thus, using the solution for uI/2 , solutions for all other i are obtained by
successively solving the tridiagonal problems at each level in reverse:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 97 / 454

Numerical Solutions of Elliptic Problems Direct Methods for Linear Systems

This results in a total of I tridiagonal problems to obtain ui , i = 0, . . . , I.


Notes:
1) Speed of FFT and cyclic reduction methods are comparable.
2) Cyclic reduction may be applied to somewhat more general equations, such
as those with variable coefficients.
3) Can accelerate by taking FFT in one direction (with constant coefficients)
and using cyclic reduction in the other.
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 98 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 99 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 100 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Returning to the difference equation (5.2) for the Poisson equation (5.1)
¯ i,j + φi−1,j + ∆
φi+1,j − 2(1 + ∆)φ ¯ (φi,j+1 + φi,j−1 ) = (∆x)2 fi,j , (5.15)
¯ = (∆x/∆y)2 .
where ∆
This may be written in general form as

Lφ = f, (5.16)

where L is the finite-difference operator for the particular problem.


Iterative methods consist of beginning with an initial guess and iteratively
“relaxing” equation (5.16) until convergence.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 101 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Jacobi Iteration
Solving equation (5.15) for φi,j and indicating the iteration number using a
superscript
1 ¯
φn+1
 n n n n
 2

i,j = ¯ φ i+1,j + φ i−1,j + ∆ φ i,j+1 + φ i,j−1 − (∆x) f i,j . (5.17)
2(1 + ∆)

Procedure:
1 Provide an initial guess φ1i,j for φi,j at each point
i = 1, . . . , I + 1, j = 1, . . . , J + 1.
2 Relax (iterate) by applying (5.17) at each grid point to produce successive
approximations:
φ2i,j , φ3i,j , . . . , φni,j , . . .
3 Continue until convergence, determined by

max |φn+1 n

i,j − φi,j |
 < ,
max |φni,j |

where, for example,  = 10−4 ⇒ convergence to ∼ 4 significant figures.


MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 102 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Notes:
1 Convergence is too slow ⇒ not used in practice.
2 Requires φn+1 n
i,j and φi,j to be stored for all i = 1, . . . , I + 1, j = 1, . . . , J + 1.
3 Used as a basis for comparison with other methods to follow.
4 Although not necessary, it is instructive to view iterative methods in matrix
form Ax = c.
→ See MMAE 501, section 2.2.2 notes.
We write A = M1 − M2 ; thus, an iterative scheme may be devised by
writing Ax = c in the form

M1 x(n+1) = M2 x(n) + c,

where we put the iteration number in parentheses in order to distinguish it


from powers. Multiplying by M−1
1 gives

x(n+1) = M−1
1 M2 x
(n)
+ M−1
1 c,

or
x(n+1) = Mx(n) + M−1
1 c,

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 103 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

where M = M−1
1 M2 is the iteration matrix.

Let
D = diagonal elements of A

−L = lower triangular elements of A

−U = upper triangular elements of A


Therefore,
Ax = (D − L − U)x = c.
Then for Jacobi iteration (M1 = D, M2 = L + U):

Dx(n+1) = (L + U)x(n) + c

x(n+1) = D−1 (L + U)x(n) + D−1 c


∴ M = M−1
1 M2 = D
−1
(L + U)
Then we check to be sure that the spectral radius of M satisfies ρ < 1 to
ensure convergence of the iterative scheme.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 104 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

In addition, a smaller spectral radius results in more rapid convergence. It


can be shown that for equation (5.17) with ∆ ¯ = 1 (∆x = ∆y) and Dirichlet
boundary conditions that
    
1 π π
ρJac (I, J) = cos + cos .
2 I +1 J +1

If I = J and I is large, then


   2
π 1 π
ρJac (I) = cos =1− + ··· ; (5.18)
I +1 2 I +1

therefore, as I → ∞, ρJac (I) → 1. As a result, we observe slower


convergence as I is increased, i.e. there is a disproportionate increase in
computational time as I is increased.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 105 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Gauss-Seidel Iteration
There are two problems with Jacobi iteration:
1 Slow convergence.
2 Must store current and previous iterations for entire grid.
The Gauss-Seidel method addresses both problems by using the most recently
updated information. For example, if sweeping along lines of constant y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 106 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Therefore, when updating φi,j , the points φi,j−1 and φi−1,j have already been
updated, and using these updated values, equation (5.17) is changed to
1 ¯
φn+1 n+1 n+1
 n n
 2

i,j = ¯ φ i+1,j + φ i−1,j + ∆ φ i,j+1 + φ i,j−1 − (∆x) f i,j . (5.19)
2(1 + ∆)

Observe that now it is not necessary to store φni,j at the previous iteration. The
values of φi,j are all stored in the same array, and it is not necessary to distinguish
between the (n) or (n + 1)st iterates. We simply use the most recently updated
information.
In matrix form, the Gauss-Seidel method is (M1 = D − L, M2 = U):

(D − L)x(n+1) = Ux(n) + c

x(n+1) = (D − L)−1 Ux(n) + (D − L)−1 c


∴ M = M−1
1 M2 = (D − L)
−1
U
It can be shown that
ρGS = ρ2Jac ;

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 107 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

therefore, for our model problem


"  2 #2  2
1 π π
ρGS (I) = ρ2Jac (I) = 1 − + ··· =1− + · · · . (5.20)
2 I +1 I +1

Thus, the rate of convergence is twice as fast as for Jacobi, i.e. the Gauss-Seidel
method requires one-half the iterations for the same level of accuracy.
Note: It can be shown that diagonal dominance of A is a sufficient (but not
necessary) condition for convergence of the Jacobi and Gauss-Seidel iteration
methods, i.e. ρ < 1, where ρ is the spectral radius of M = M−1
1 M2 (see Morton
and Mayers, p. 205 for proof).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 108 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Successive Overrelaxation (SOR)


In Gauss-Seidel iteration, the sign of the error typically does not change from
iteration to iteration; therefore, the iterative solution normally approaches the
exact solution very slowly. Convergence can often be accelerated by
“over-relaxing,” or magnifying, the change at each iteration.
This is accomplished by taking a weighted average of the previous iterate φni,j and
the Gauss-Seidel iterate φn+1
i,j . So if we denote the Gauss-Seidel iterate (5.19) by

φi,j , the new SOR iterate is given by

φn+1 n
i,j = (1 − ω)φi,j + ωφi,j , (5.21)

where ω is the relaxation parameter and 0 < ω < 2 for convergence (Morton &
Mayers, p. 206).
ω=1 ⇒ Gauss-Seidel

1<ω<2 ⇒ Overrelaxation

0<ω<1 ⇒ Underrelaxation

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 109 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Overrelaxation typically accelerates convergence of linear problems, while


underrelaxation is often required for convergence when dealing with non-linear
problems.
In matrix form, the SOR method is (M1 = D − ωL, M2 = (1 − ω)D + ωU):

(D − ωL)x(n+1) = [(1 − ω)D + ωU] x(n) + ωc

x(n+1) = (D − ωL)−1 [(1 − ω)D + ωU] x(n) + ω(D − ωL)−1 c


∴ M = M−1
1 M2 = (D − ωL)
−1
[(1 − ω)D + ωU]
It can be shown that for the Poisson equation, the optimal value of ω is
2
ωopt = p , (5.22)
1 + 1 − ρ2Jac

and for this ωopt , the spectral radius for SOR is


!2
ρ
ρSOR = pJac . (5.23)
1 + 1 − ρ2Jac

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 110 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

 2
1 π
Recall that for our model problem, ρJac = 1 − 2 I+1 + · · · ; thus,

2
ωopt = p
1+ 1 − ρ2Jac
2
= s  2
 2
1 π
1+ 1− 1− 2 I+1 + ···

2 (5.24)
= s  
 2
π
1+ 1 − 1 − I+1 + · · ·

2
ωopt ≈ π .
1 + I+1

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 111 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Then for large I


  2 2
1 π
1 − 2 I+1
ρSOR ≈

π
1+
 
I+1
("  2 #  )2
1 π π
≈ 1− 1− + ···
2 I +1 I +1 (5.25)
 2
π
≈ 1− + ···
I +1

ρSOR ≈ 1− .
I +1

Thus, as I → ∞, ωopt → 2, and ρSOR → 1.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 112 / 454
Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

However, from a comparison of equations (5.25) and (5.20), we see that

ρSOR < ρGS ,

such that SOR converges at a rate 2(I+1)


π times faster than Gauss-Seidel if the
optimal value of the relaxation parameter is used for the model problem under
consideration. Therefore, the relative convergence rate improves linearly with
increasing I.
This analysis assumes that we know ωopt . Typically, we do not, and the rate of
convergence depends significantly on the choice of ω, e.g. the typical behavior for
linear problems is as follows (for given I):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 113 / 454

Numerical Solutions of Elliptic Problems Iterative (Relaxation) Methods

Notes:
1 Although ωopt does not depend on the right-hand-side, it does depend upon:
i) differential equation
ii) method of discretization
iii) boundary conditions
iv) shape of domain
2 For a given problem, ωopt must be estimated from a similar problem and/or
trial and error.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 114 / 454
Numerical Solutions of Elliptic Problems Boundary Conditions

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 115 / 454

Numerical Solutions of Elliptic Problems Boundary Conditions

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 116 / 454
Numerical Solutions of Elliptic Problems Boundary Conditions

Dirichlet Boundary Conditions


Simply apply the specified values at the boundaries to φ1,j , φI+1,j , φi,1 , and
φi,J+1 and iterate on the Jacobi (5.17), Gauss-Seidel (5.19), or SOR (5.21)
equation in the interior, i = 2, . . . , I, j = 2, . . . , J.
For example, Jacobi or Gauss-Seidel iterate at i = 2, j = 2:

1 ¯
φn+1
 n n
 2

2,2 = ¯ φ 3,2 + φ 1,2 + ∆ φ 2,3 + φ 2,1 − (∆x) f 2,2 .
2(1 + ∆)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 117 / 454

Numerical Solutions of Elliptic Problems Boundary Conditions

Neumann (Derivative) Boundary Conditions


Consider the following boundary condition
∂φ
= c at x = 0. (5.26)
∂x

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 118 / 454
Numerical Solutions of Elliptic Problems Boundary Conditions

The simplest treatment would be to use the Jacobi (5.17), Gauss-Seidel (5.19), or
SOR (5.21) equation to update φi,j in the interior for i = 2, . . . , I, and then to
approximate the boundary condition (5.26) by a forward difference applied at i = 1

φ2,j − φ1,j
+ O(∆x) = c. (5.27)
∆x
This could then be used to update φ1,j , j = 2, . . . , J using

φ1,j = φ2,j − c∆x.

This is unacceptable, however, because (5.27) is only first-order accurate.


A better alternative is to update the interior points as before, but now apply the
difference equation at the boundary. For example, we could apply Jacobi (5.17) at
i=1
1 ¯
φn+1
 n n n n
 2

1,j = ¯ φ 2,j + φ 0,j + ∆ φ 1,j+1 + φ 1,j−1 − (∆x) f 1,j . (5.28)
2(1 + ∆)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 119 / 454

Numerical Solutions of Elliptic Problems Boundary Conditions

However, this involves a value φn0,j that is outside the domain. A second-order
accurate central-difference approximation for the boundary condition (5.26) is

φn2,j − φn0,j
+ O(∆x2 ) = c, (5.29)
2∆x
which also involves the value φn0,j . Therefore, solving (5.29) for φn0,j gives

φn0,j = φn2,j − 2c∆x,

and substituting into the difference equation (5.28) to eliminate φn0,j leads to

1 ¯
φn+1 n n n 2
   
1,j = ¯ 2 φ 2,j − c∆x + ∆ φ 1,j+1 + φ 1,j−1 − (∆x) f1,j . (5.30)
2(1 + ∆)

Thus, we use (5.30) to update φn+1


1,j , j = 2, . . . , J.

Notes:
1 This is the same procedure used for a Dirichlet condition but with an
additional sweep along the left boundary using (5.30) for φn+1
1,j , j = 2, . . . , J.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 120 / 454
Numerical Solutions of Elliptic Problems Boundary Conditions

2 This approach requires special treatment at corners depending upon the


boundary condition along the adjacent boundary. For example, consider if in
addition to (5.26) we have

∂φ
=d at y = b. (5.31)
∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 121 / 454

Numerical Solutions of Elliptic Problems Boundary Conditions

Applying (5.30) at the corner i = 1, j = J + 1


1 ¯ φn
φn+1 n n 2
   
1,J+1 = ¯ 2 φ 2,J+1 − c∆x + ∆ 1,J+2 + φ 1,J − (∆x) f1,J+1 ,
2(1 + ∆)
(5.32)
n
where φ1,J+2 is outside the domain. Approximating (5.31) using a central
difference in the same manner as (5.29) gives
φn1,J+2 − φ
d,
=
which leads to
φn1,J+2 = φn1,J + 2d∆y.
Substituting into (5.32) to eliminate φn1,J+2 gives
1 ¯
φn+1 n n 2
   
1,J+1 = ¯ 2 φ 2,J+1 − c∆x + 2 ∆ φ 1,J + d∆y − (∆x) f1,J+1 ,
2(1 + ∆)
which is used to update the corner value i = 1, j = J + 1.
3 If a lower-order approximation, such as (5.27), is used at a boundary, its
truncation error generally dominates the convergence rate.
4 The above approach is used in thomas.f, etc.
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 122 / 454
Numerical Solutions of Elliptic Problems Alternating-Direction-Implicit (ADI) Method

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 123 / 454

Numerical Solutions of Elliptic Problems Alternating-Direction-Implicit (ADI) Method

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 124 / 454
Numerical Solutions of Elliptic Problems Alternating-Direction-Implicit (ADI) Method

Recall that in elliptic problems, the solution anywhere depends on the solution
everywhere, i.e. it has an infinite speed of propagation in all directions.
However, in Jacobi, Gauss-Seidel, and SOR, information only propagates through
the mesh one point at a time. For example, if sweeping along lines of constant y
with 0 ≤ x ≤ a, it takes I iterations before the boundary condition at x = a is
“felt” at x = 0.
⇒ These techniques are not very “elliptic like.”
A more “elliptic-like” method could be obtained by solving entire lines in the grid
in an implicit manner. For example, sweeping along lines of constant y and
solving each constant y-line implicitly, i.e. all at once, would allow for the
boundary condition at x = a to influence the solution in the entire domain after
only one sweep through the grid.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 125 / 454

Numerical Solutions of Elliptic Problems Alternating-Direction-Implicit (ADI) Method

Returning to the difference equation (5.15) for the Poisson equation


¯ i,j + φi−1,j + ∆
φi+1,j − 2(1 + ∆)φ ¯ (φi,j+1 + φi,j−1 ) = (∆x)2 fi,j . (5.33)

Consider the j th line and assume that values along the j + 1st and j − 1st lines
are taken from the previous iterate.
Rewriting (5.33) as an implicit equation for the values of φi,j along the j th line
gives

φn+1 ¯ n+1 n+1 2 ¯ n n



i+1,j − 2(1 + ∆)φi,j + φi−1,j = (∆x) fi,j − ∆ φi,j+1 + φi,j−1 , i = 2, . . . , I.
(5.34)
th
Therefore, we have a tridiagonal problem for φi,j along the j line, which can be
solved using the Thomas algorithm.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 126 / 454
Numerical Solutions of Elliptic Problems Alternating-Direction-Implicit (ADI) Method

Notes:
1 If sweeping through j-lines, j = 2, . . . , J, then φni,j−1 becomes φn+1
i,j−1 in
(5.34), i.e. it has already been updated. Therefore, we use updated values as
in Gauss-Seidel.
2 Can also incorporate SOR.
3 More efficient at spreading information throughout the domain; therefore, it
reduces the number of iterations required for convergence, but there is more
computation per iteration.
4 This provides motivation for the ADI method.

ADI Method
In the ADI method we sweep along lines but in alternating directions.
In the first half of the iteration we perform a sweep along constant y-lines by
solving the series of tridiagonal problems for j = 2, . . . , J:
n+1/2 n+1/2 2 n+1/2
φi+1,j −h(2 + σ)φi,j + φi−1,j = (∆x) i fi,j
−∆ ¯ φn σ
 n n+1/2 (5.35)
i,j+1 − 2 − ∆
¯ φi,j + φi,j−1 , i = 2, . . . , I.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 127 / 454

Numerical Solutions of Elliptic Problems Alternating-Direction-Implicit (ADI) Method

Notes:
n+1/2
1 The φi,j−1 term on the right-hand-side has been updated from the previous
line.
2 Unlike in equation (5.34), differencing in the x- and y-directions are kept
separate to mimic diffusion in each direction. This is called a splitting
method.
3 σ is an acceleration parameter to enhance diagonal dominance (σ > 0).
σ = 0 corresponds to no acceleration. Note that the σ terms on each side of
the equation cancel.
In the second half of the iteration we sweep along constant x-lines by solving the
series of tridiagonal problems for i = 2, . . . , I:
¯ n+1 − (2∆
∆φ ¯ + σ)φn+1 + ∆φ
¯ n+1 = (∆x)2 fi,j
i,j+1
h i,j i,j−1 i
n+1/2
− φi+1,j − (2 − σ) φi,j
n+1/2
+ φn+1 (5.36)
i−1,j , j = 2, . . . , J,

where φn+1
i−1,j has been updated from the previous line.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 128 / 454
Numerical Solutions of Elliptic Problems Alternating-Direction-Implicit (ADI) Method

Notes:
1 Involves (I − 1) + (J − 1) tridiagonal solves for each iteration (for Dirichlet
boundary conditions).
2 For ∆x = ∆y (∆ ¯ = 1), it can be shown that for the Poisson (or Laplace)
equation with Dirichlet boundary conditions that the acceleration parameter
that gives the best speedup is

σ = 2 sin (π/R) ,

where R = max(I + 1, J + 1).


3 The ADI method with splitting is typically faster than Gauss-Seidel or SOR
for equations in which the terms are easily factored into x and y directions,
but it is more difficult if have derivative mixed terms such as ∂ 2 φ/∂x∂y.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 129 / 454

Numerical Solutions of Elliptic Problems Compact Finite Differences

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 130 / 454
Numerical Solutions of Elliptic Problems Compact Finite Differences

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 131 / 454

Numerical Solutions of Elliptic Problems Compact Finite Differences

As we have seen before, in order to obtain higher-order finite-difference


approximations it is necessary to incorporate more points in the finite-difference
stencil. However, it is advantageous to maintain the compact nature of
second-order accurate central differences because they produce systems of
equations, e.g. tridiagonal, that may be solved efficiently. Here we derive a
compact fourth-order accurate central-difference approximation for the
second-order derivatives in the Poisson equation.
For convenience, we define the following second-order accurate central-difference
operators
φi−1,j − 2φi,j + φi+1,j
δx2 φi,j = ,
∆x2
φi,j−1 − 2φi,j + φi,j+1
δy2 φi,j = .
∆y 2
Recall from equation (3.13) in section 3.2 that from the Taylor series, we have

∂2φ 2 ∆x2 ∂ 4 φ
= δx φ − + O(∆x4 ). (5.37)
∂x2 12 ∂x4

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 132 / 454
Numerical Solutions of Elliptic Problems Compact Finite Differences

Therefore,
∂ 2 φ ∆x2 ∂ 4 φ
δx2 φ = 2
+ 4
+ O(∆x4 ).
∂x 12 ∂x
(5.38)
∆x2 ∂ 2 ∂ 2 φ
 
δx2 φ = 1+ 2 2
+ O(∆x4 ).
12 ∂x ∂x
But from equation (5.37)
∂2
= δx2 + O(∆x2 ).
∂x2
Substituting into (5.38) gives
2
  2
∆x 2 2 ∂ φ
+ O(∆x4 ),

δx2 φ = 1+ δx + O(∆x ) 2
12 ∂x
∆x2 2 ∂ 2 φ
 
δx2 φ = 1+ δ + O(∆x4 ).
12 x ∂x2

Solving for ∂ 2 φ/∂x2


−1 −1
∂2φ ∆x2 2 2
 
∆x
= 1+ δ δx2 φ + 1 + δ2 O(∆x4 ) (5.39)
∂x2 12 x 12 x
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 133 / 454

Numerical Solutions of Elliptic Problems Compact Finite Differences

From a binomial expansion (with ∆x sufficiently small)


−1
∆x2 2 ∆x2 2

1+ δ =1− δ + O(∆x4 ). (5.40)
12 x 12 x

Because the last term in equation (5.39) is still O(∆x4 ), we can write equation
(5.39) as
−1
∂2φ ∆x2 2

2
= 1+ δx δx2 φ + O(∆x4 ). (5.41)
∂x 12
Substituting the expression (5.40) into equation (5.41) leads to a O(∆x4 )
accurate central-difference approximation for the second derivative

∂2φ ∆x2 2 2
 
2
= 1− δx δx φ + O(∆x4 ).
∂x 12

Due to the δx2 (δx2 φ) operator, however, this approximation involves the five points
φi−2 , φi−1 , φi , φi+1 and φi+2 ; therefore, it is not compact.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 134 / 454
Numerical Solutions of Elliptic Problems Compact Finite Differences

In order to obtain a compact scheme, we also consider the derivative in the


y-direction. Similar to equation (5.41), we have in the y-direction
−1
∂2φ ∆y 2 2

= 1+ δ δy2 φ + O(∆y 4 ). (5.42)
∂y 2 12 y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 135 / 454

Numerical Solutions of Elliptic Problems Compact Finite Differences

Now consider the Poisson equation

∂2φ ∂2φ
+ 2 = f (x, y).
∂x2 ∂y

Substituting equations (5.41) and (5.42) into the Poisson equation leads to
−1 −1
∆x2 2 ∆y 2 2
 
1+ δ
+ 1+ δy δx2 φ
δy2 φ + O(∆4 ) = f (x, y),
12 x 12
  
2 ∆y 2 2
where ∆ = max(∆x, ∆y). Multiplying by 1 + ∆x12 x δ 2
1 + 12 y gives
δ

∆y 2 2 2 ∆x2 2 2
   
1+ δy δx φ + 1 + δx δy φ + O(∆4 )
12 12
(5.43)
∆x2 2 ∆y 2 2
 
4
= 1+ δ + δ + O(∆ ) f (x, y),
12 x 12 y

which is a fourth-order accurate finite-difference approximation to the Poisson


equation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 136 / 454
Numerical Solutions of Elliptic Problems Compact Finite Differences

Expand the first term in equation (5.43) as follows

∆y 2 2 2 ∆y 2 2 φi−1,j − 2φi,j + φi+1,j


   
1+ δ δ φ = 1+ δ
12 y x 12 y ∆x2
1
= (φi−1,j − 2φi,j + φi+1,j )
∆x2
1
+ [(φi−1,j−1 − 2φi−1,j + φi−1,j+1 )
12∆x2
−2 (φi,j−1 − 2φi,j + φi,j+1 )
+ (φi+1,j−1 − 2φi+1,j + φi+1,j+1 )]
1
= [−20φi,j + 10 (φi−1,j + φi+1,j )
12∆x2
−2 (φi,j−1 + φi,j+1 )
+φi−1,j−1 + φi−1,j+1 + φi+1,j−1 + φi+1,j+1 ] .

Therefore, we have a nine-point stencil, but the approximation only requires three
points in each direction and thus it is compact.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 137 / 454

Numerical Solutions of Elliptic Problems Compact Finite Differences

Similarly, expanding the second term in equation (5.43) yields

∆x2 2 2
 
1
1+ δx δy φ = [−20φi,j + 10 (φi,j−1 + φi,j+1 )
12 12∆y 2
−2 (φi−1,j + φi+1,j ) + φi−1,j−1 + φi−1,j+1 + φi+1,j−1 + φi+1,j+1 ] ,

and the right-hand-side of equation (5.43) is

∆x2 2 ∆y 2 2
 
1
1+ δx + δy f (x, y) = fi,j + [fi−1,j − 2fi,j + fi+1,j
12 12 12
+fi,j−1 − 2fi,j + fi,j+1 ]
1
= [8fi,j + fi−1,j + fi+1,j + fi,j−1 + fi,j+1 ] .
12

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 138 / 454
Numerical Solutions of Elliptic Problems Compact Finite Differences

Thus, the coefficients in the finite-difference stencil for φ(x, y) are as follows:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 139 / 454

Numerical Solutions of Elliptic Problems Compact Finite Differences

and for f (x, y) they are

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 140 / 454
Numerical Solutions of Elliptic Problems Compact Finite Differences

Notes:
1 Observe that in equation (5.43) the two-dimensionality of the equation has
been taken advantage of to obtain the compact finite-difference stencil, i.e.
see the δx2 δy2 φ and δy2 δx2 φ difference operators.
2 Because the finite-difference stencil is compact, i.e. only involving three
points in each direction, application of the ADI method as in the previous
section results in a set of tridiagonal problems to solve. Therefore, this
fourth-order, compact finite-difference approach is no less efficient then that
for the second-order scheme used in the previous section (there are simply
additional terms on the right-hand-side of the equations).
3 The primary disadvantage of using higher-order schemes is that it is generally
necessary to use lower-order approximations for derivative boundary
conditions. This is not a problem, however, for Dirichlet boundary conditions.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 141 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 142 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 143 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Motivation
The iterative techniques we have discussed all have the following property:
High-frequency components of error ⇒ fast convergence.
Low-frequency components of error ⇒ slow convergence.
To illustrate this, consider the simple one-dimensional problem

d2 φ
= 0, 0 ≤ x ≤ 1, (5.44)
dx2
with φ(0) = φ(1) = 0, which has the exact solution φ(x) = 0. Therefore, all plots
of the numerical solution are also plots of the error.
Discretizing equation (5.44) at I + 1 points using central differences gives

φi+1 − 2φi + φi−1 = 0, i = 1, . . . , I − 1


φ0 = φI = 0.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 144 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

To show how the nature of the error affects convergence, consider an initial guess
consisting of the Fourier mode φ(x) = sin(kπx), where k is the wavenumber and
indicates the number of half sine waves on the interval 0 ≤ x ≤ 1. In discretized
form, with xi = i∆x = i/I, this is
 
kπi
φi = sin , i = 0, . . . , I,
I

with the wavenumber 1 ≤ k ≤ I − 1. Thus,


Small k ⇒ long, smooth waves (low frequency).
Large k ⇒ highly oscillatory waves (high frequency).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 145 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Consider initial guesses with k = 1, 3, 6 (the figures are from Briggs et al.):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 146 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Applying Jacobi and Gauss-Seidel iteration with I = 64, the solution converges
more rapidly for the higher frequency initial guess.
Jacobi:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 147 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Gauss-Seidel:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 148 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

A more realistic situation is one in which the initial guess contains multiple modes,
for example
      
1 πi 6πi 32πi
φi = sin + sin + sin .
3 I I I

The k = 1, 6, and 32 terms represent low-, medium- and high-frequency modes,


respectively.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 149 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Applying Jacobi iteration with I = 64, the error is reduced rapidly during the early
iterations but more slowly thereafter.

Thus, there is rapid convergence until the high-frequency modes are smoothed
out, then slow convergence for the lower frequency modes.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 150 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

See, for example, the following sequence using Jacobi iteration:


k = 3 after 1 iteration (left) and 10 iterations (right):

k = 16 after 1 iteration (left) and 10 iterations (right):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 151 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

k = 2 and k = 16 after 1 iteration (left) and 10 iterations (right):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 152 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Multigrid methods take advantage of this property of relaxation techniques by


recognizing that smooth components of the error become more oscillatory with
respect to the grid size on a coarse grid.

Thus, relaxation is more effective on a coarse grid representation of the error (it is
also faster).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 153 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Notes:
1) Multigrid methods are not so much a specific set of techniques as they are a
framework for accelerating relaxation (iterative) methods.
2) Multigrid methods are comparable in speed with fast direct methods, such as
Fourier methods and cyclic reduction, but can be used to solve general
elliptic equations with variable coefficients and even nonlinear equations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 154 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Multigrid Methodology
Consider the general second-order linear elliptic partial differential equation of the
form
∂2φ ∂φ ∂2φ ∂φ
A(x, y) 2 +B(x, y) +C(x, y) 2 +D(x, y) +E(x, y)φ = F (x, y). (5.45)
∂x ∂x ∂y ∂y

To be elliptic, A(x, y)C(x, y) > 0 for all (x, y). Approximating this differential
equation using second-order accurate central differences gives
φi+1,j − 2φi,j + φi−1,j φi+1,j − φi−1,j
Ai,j 2
+ Bi,j
∆x 2∆x
φi,j+1 − 2φi,j + φi,j−1 φi,j+1 − φi,j−1
Ci,j 2
+ Di,j
∆y 2∆y
+Ei,j φi,j = Fi,j ,

where Ai,j = A(xi , yj ), etc. We rewrite this equation in the form

ai,j φi+1,j + bi,j φi−1,j + ci,j φi,j+1 + di,j φi,j−1 + ei,j φi,j = Fi,j , (5.46)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 155 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

where
Ai,j Bi,j Ai,j Bi,j
ai,j = + , bi,j = − ,
∆x2 2∆x ∆x2 2∆x
Ci,j Di,j Ci,j Di,j
ci,j = + , di,j = − ,
∆y 2 2∆y ∆y 2 2∆y
2Ai,j 2Ci,j
ei,j = Ei,j − − .
∆x2 ∆y 2
For convenience, write (5.46) (or some other difference equation) as

Lφ = f, (5.47)

where L represents the difference operator. If φ is the exact solution to equation


(5.46), then the error is defined by

e = φ − φ̄, (5.48)

where φ̄ is an approximation to φ. The residual is defined by

r = f − Lφ̄. (5.49)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 156 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Observe from (5.47) that if φ̄ = φ, then the residual is zero; therefore, the residual
is a measure of how “wrong” the approximate solution is.
Substituting (5.48) into equation (5.47) gives the error equation
Le = r = f − Lφ̄. (5.50)
From these definitions, we can devise a scheme with which to correct the solution
on a fine grid by solving for the error on a coarse grid.
⇒ Coarse-Grid Correction (CGC)
Definitions:
Fine grid → Ωh ; Coarse grid → Ω2h

→ Grids are typically reduced by a factor of two in each direction.


MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 157 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Restriction operator: Ih2h


→ Move information from fine grid to coarse grid.
h
Interpolation (Prolongation) operator: I2h
→ Move information from coarse grid to fine grid.
Coarse-Grid Correction Sequence:
1) Relax Lh φ̄h = f h using Gauss-Seidel, ADI, etc. ν1 times on the fine grid Ωh
with an initial guess φ̄h .
2) Compute the residual on Ωh and restrict it to the coarse grid Ω2h :

r2h = Ih2h rh = Ih2h (f h − Lh φ̄h ).

3) Using the residual as the right-hand-side, ’solve’ the error equation


L2h e2h = r2h on the coarse grid Ω2h .
4) Interpolate the error to the fine grid and correct the fine-grid approximation
according to
φ̄h ← φ̄h + I2h
h 2h
e .
5) Relax Lh φ̄h = f h ν2 times on the fine grid Ωh with corrected approximation
φ̄h as the initial guess.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 158 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Note: In practice, ν1 and ν2 are small (1, 2, or 3).

This CGC scheme is the primary component of all the many multigrid algorithms.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 159 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

To illustrate consider the CGC sequence for equation (5.44), i.e. φ00 = 0
(I = 64, ν1 = 3):
Initial guess:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 160 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Step 1 after one relaxation sweep on fine grid:

→ The high-frequency mode is eliminated after only one relaxation step.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 161 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Step 1 after three relaxation sweeps on fine grid:

→ The low-frequency mode is reduced very little after three relaxation step.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 162 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Step 3 after one relaxation sweep on coarser grid:

→ ... and not much more after three.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 163 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Step 3 after three relaxation sweeps on coarser grid:

→ Restriction to the coarser grid accelerates convergence of the


low-frequency mode, which has higher frequency relative to the coarser grid.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 164 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Step 5 after three relaxation sweeps on fine grid:

→ Low-frequency mode is nearly eliminated after CGC.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 165 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

How do we obtain the coarse grid solution for e2h in step 3?


→ Recursively replace step 3 by additional CGCs on progressively coarser grids.
⇒ V-Cycle

For example, a four-grid V-cycle is illustrated by

The V-cycles are repeated until convergence.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 166 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Note: The error being solved for on successively coarser grids is the error of the
error on the next finer grid, i.e. on grid...
Ωh → relaxation on original equation for φ.
Ω2h → relaxation on equation for error on Ωh .
Ω4h → relaxation on equation for error on Ω2h .
..
.
This simple V-cycle scheme is appropriate when a good initial guess is available.
For example, when considering a solution to equation (5.45) in the context of an
unsteady calculation in which case the solution for φh from the previous time step
is a good initial guess for the current time step.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 167 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

If no good initial guess is available, then Full Multigrid V-Cycle (FMV) may be
applied according to the following procedure:
1) Solve Lφ = f on the coarset grid (Note: φ not e)
2) Interpolate to next finer grid.
3) Perform V-cycle to correct φ.
4) Interpolate to next finer grid.
5) Repeat (3) and (4) until finest grid is reached.
6) Perform V-cycles until convergence.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 168 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Grid Definitions:
Because each successive grid differs by a factor of two, the finest grid size is often
taken as 2n + 1, where n is an integer. Somewhat more general grids may be
obtained using the following grid definitions.
The differential equation (5.45) is discretized on a uniform grid having Nx × Ny
points which are defined by

Nx = mx 2(nx −1) + 1, Ny = my 2(ny −1) + 1, (5.51)

where nx and ny determine the number of grid levels, and mx and my determine
the size of the coarsest grid, which is (mx + 1) × (my + 1).
For a given grid, nx and ny should be as large as possible, and mx and my should
be as small as possible for maximum efficiency (typically mx and my are 2, 3 or
5). For example:
Nx = 65 ⇒ mx = 2, nx = 6
Nx = 129 ⇒ mx = 2, nx = 7
Nx = 49 ⇒ mx = 3, nx = 5
Nx = 81 ⇒ mx = 5, nx = 5
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 169 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

The number of grid levels is given by

N = max(nx , ny ) (5.52)

resulting in the series of grids

G(1) < . . . < G(L) < . . . < G(N ), (5.53)

where G(1) is the coarsest grid, G(N ) is the finest grid and L = 1, . . . , N . Each
grid G(L) has Mx (L) × My (L) grid points, where

Mx (L) = mx 2[max(nx +L−N,1)−1] + 1,


(5.54)
My (L) = my 2[max(ny +L−N,1)−1] + 1.

For example,

Nx = 65, Ny = 49 ⇒ mx = 2, nx = 6 and my = 3, ny = 5.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 170 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Thus, N = max(nx , ny ) = 6 and

G(6) : Mx (6) = 65, My (6) = 49


G(5) : Mx (5) = 33, My (5) = 25
G(4) : Mx (4) = 17, My (4) = 13
G(3) : Mx (3) = 9, My (3) = 7
G(2) : Mx (2) = 5, My (2) = 4
G(1) : Mx (1) = 3, My (1) = 4

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 171 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Boundary Conditions
At each boundary the general form of the boundary condition is
∂φ
pφ + q = s, (5.55)
∂n
where n is the direction normal to the surface. This boundary condition is applied
directly on the finest grid Ωh , i.e.
h
h h h ∂φ
p φ +q = sh . (5.56)
∂n
But on the coarser grids, we need the boundary condition for the error. In order to
obtain such a condition, consider the following. On the coarse grid Ω2h , equation
(5.55) applies to the solution φ; thus,
2h
2h 2h 2h ∂φ
p φ +q = s2h .
∂n

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 172 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

If we enforce the error to be zero on the boundaries, then from


e2h = φ2h − φ̄2h = 0, i.e. φ2h = φ̄2h , we also have

∂ φ̄2h
p2h φ̄2h + q 2h = s2h .
∂n
Therefore, to obtain a boundary condition for the error on Ω2h
2h 2h 2h
 
2h 2h 2h ∂e 2h 2h 2h ∂φ 2h 2h 2h ∂ φ̄
p e +q = p φ +q − p φ̄ + q
∂n ∂n ∂n
= s2h − s2h (5.57)
∂e2h
p2h e2h + q 2h = 0.
∂n
Thus, the boundary conditions are homogeneous on all but the finest grid, where
the original condition on φ is applied. For example,
Dirichlet ⇒ e = 0 on boundary.
∂e
Neumann ⇒ ∂n = 0 on boundary.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 173 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

That is, the type of boundary condition (Dirichlet, Neumann, or Robin) does not
change, i.e. the p and q coefficients are the same, but they become homogeneous,
i.e. s = 0.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 174 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Relaxation
Typically, red-black Gauss-Seidel iteration is used to relax the difference equation:

By performing the relaxation on all of the red and black points separately, it
eliminates data dependencies such that it is easily implemented on parallel
computers (see section 10). Note that when Gauss-Seidel is used, SOR should not
be implemented because it destroys the high-frequency smoothing.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 175 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

The point-by-point relaxation of Gauss-Seidel does not work well, however, if


there are large gradients in the solution or rapid variations in the coefficients in
the differential equation (5.45). In such cases it is better to use line relaxation for
the same reason that ADI is better than Gauss-Seidel.
In general, we may have such behavior in both directions; therefore, we could use
alternating-direction-line (ADL) relaxation. When sweeping along lines of
constant y, the following tridiagonal problem is solved for each j = 1, . . . , My (L)
(see equation (5.46))

ai,j φi+1,j + ei,j φi,j + bi,j φi−1,j = fi,j − ci,j φ∗i,j+1 − di,j φ∗i,j−1 , (5.58)

for i = 1, . . . , Mx (L). Here φ∗ denotes the most recent approximation. Similar to


red-black Gauss-Seidel, we could sweep all lines with j even and j odd separately.
We will refer to this as zebra relaxation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 176 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Then lines of constant x are swept by solving the tridiagonal problem for each
i = 1, . . . , Mx (L) given by

ci,j φi,j+1 + ei,j φi,j + di,j φi,j−1 = fi,j − ai,j φ∗i+1,j − bi,j φ∗i−1,j , (5.59)

for j = 1, . . . , My (L). Again we could sweep all lines with i even and i odd
separately.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 177 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Restriction Operator: Ih2h


The restriction operator is required for moving information from the finer grid to
the coarser grid.
Notation:
Grid Indices Range
Coarser (i, j) 1 ≤ i ≤ N XC, 1 ≤ j ≤ N Y C
Finer (i∗ , j ∗ ) 1 ≤ i∗ ≤ N XF, 1 ≤ j ∗ ≤ N Y F
Applying restriction in both directions:

N XC = 12 (N XF − 1) + 1, N Y C = 12 (N Y F − 1) + 1
i∗ = 2i − 1, j ∗ = 2j − 1

The easiest restriction operator is straight injection for which

φ2h h
i,j = φi∗ ,j ∗ ,

i.e. we simply drop the points that are not common to both grids. The matrix
symbol for straight injection is [1].

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 178 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

A better operator is full weighting. The matrix symbol is

Thus,
1
 
φ2h
i,j = h h h h
16 φi∗ −1,j ∗ −1 + φi∗ −1,j ∗ +1 + φi∗ +1,j ∗ −1 + φi∗ +1,j ∗ +1
+ 18 φhi∗ ,j ∗ −1 + φhi∗ ,j ∗ +1 + φhi∗ −1,j ∗ + φhi∗ +1,j ∗
 
(5.60)
+ 14 φhi∗ ,j ∗

This represents a weighted average of surrounding points in the fine mesh.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 179 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

We then use straight injection on the boundaries, i.e. φ2h h


i,j = φi∗ ,j ∗ ,
i = 1, N XC, j = 1, . . . , N Y C and j = 1, N Y C, i = 1, . . . , N XC.
If, for example, restriction is applied only in the x-direction, then N Y C = N Y F
and j ∗ = j in equation (5.60).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 180 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

h
Interpolation (Prolongation) Operator: I2h
The interpolation operator is required for moving information from the coarser to
finer grid. The most commonly used interpolation operator is based on bilinear
interpolation.

φhi∗ ,j ∗ = φ2h
i,j ← copy common points
1

φhi∗ +1,j ∗ = 2 φ2h
i,j + φ 2h
i+1,j
1

φhi∗ ,j ∗ +1 = 2 φ2h 2h
i,j + φi,j+1
1

φhi∗ +1,j ∗ +1 = 4 φ2h 2h 2h 2h
i,j + φi+1,j + φi,j+1 + φi+1,j+1
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 181 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

Speed Comparisons
Consider the test problem

∂2φ ∂φ ∂2φ ∂φ
A(x) + B(x) + C(y) + D(y) = F (x, y),
∂x2 ∂x ∂y 2 ∂y
with Neumann boundary conditions. The following times are for an SGI Indy
R5000-150MHz. The grid is N × N .
ADI:

 = 10−4  = 10−5
N Iterations Time (sec) Iterations Time (sec)
65 673 22.35 821 27.22
129 2, 408 366.06 2, 995 456.03

Note that in both cases, the total time required for the N = 129 case is
approximately 16× that with N = 65 (∼ 4× increase in points and ∼ 4× increase
in iterations).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 182 / 454
Numerical Solutions of Elliptic Problems Multigrid Methods

Multigrid:
V-cycle with ADL relaxation (no FMV to get improved initial guess). Here the
convergence criterion  is evaluated between V-cycles.

 = 10−4  = 10−5
N Iterations Time (sec) Iterations Time (sec)
65 18 1.78 23 2.28
129 23 10.10 29 12.68

Notes:
1) In both cases, the total time required for the N = 129 case is approximately
6× that with N = 65 (the minimum is 4×).
⇒ The multigrid method scales to larger grid sizes more effectively than ADI
alone, i.e. note the small increase in the number of V-cycles with increasing N .
2) The case with N = 65 is approximately 13× faster than ADI, and the case
with N = 129 is approximately 36× faster!

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 183 / 454

Numerical Solutions of Elliptic Problems Multigrid Methods

3) References:
1 Briggs, W.C., Henson, V.E. and McCormick, S.F., “A Multigrid Tutorial,”
(2nd Edition) SIAM (2000).
2 Thomas, J.L., Diskin, B. and Brandt, A.T., “Textbook Multigrid Efficiency for
Fluid Simulations,” Ann. Rev. Fluid Mech. (2003), 35, pp. 317–340.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 184 / 454
Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Outline

5 Numerical Solutions of Elliptic Problems


Introduction
Finite-Difference Methods for the Poisson Equation
Direct Methods for Linear Systems
Fourier Transform Methods
Cyclic Reduction
Iterative (Relaxation) Methods
Jacobi Iteration
Gauss-Seidel Iteration
Successive Overrelaxation (SOR)
Boundary Conditions
Dirichlet Boundary Conditions
Neumann (Derivative) Boundary Conditions
Alternating-Direction-Implicit (ADI) Method
Compact Finite Differences
Multigrid Methods
Motivation
Multigrid Methodology
Speed Comparisons

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 185 / 454

Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Outline (cont’d)
Treatment of Nonlinear Convective Terms
Upwind-Downwind Differencing

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 186 / 454
Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Consider the 2-D, steady Burger’s equations

∂2u ∂2u
 
∂u ∂u
Re u +v = + 2, (5.61)
∂x ∂y ∂x2 ∂y

∂2v ∂2v
 
∂v ∂v
Re u +v = + 2, (5.62)
∂x ∂y ∂x2 ∂y
which represent a simplified version of the Navier-Stokes equations as there are no
pressure terms. The terms on the left-hand-side are the convection terms, and
those on the right-hand-side are the viscous or diffusion terms.
The Burger’s equations are elliptic due to the nature of the second-order viscous
terms, but the convection terms make the equations non-linear.
A simple approach to linearizing the equations is known as Picard iteration in
which we take the coefficients of the non-linear (first derivative) terms to be
known from the previous iteration denoted by

u∗i,j , vi,j

.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 187 / 454

Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Let us begin by approximating (5.61) using central differences for all derivatives as
follows
 
∗ ui+1,j − ui−1,j ∗ ui,j+1 − ui,j−1
Re ui,j + vi,j
2∆x 2∆y
ui+1,j − 2ui,j + ui−1,j ui,j+1 − 2ui,j + ui,j−1
= 2
+ .
(∆x) (∆y)2

Multiplying by (∆x)2 and rearranging leads to

1 − 21 Re ∆x u∗i,j ui+1,j + 1 + 12 Re ∆x u∗i,j ui−1,j


 

¯ − 1 Re ∆x ∆ ¯ 1/2 v ∗ ui,j+1 + ∆
¯ + 1 Re ∆x ∆¯ 1/2 v ∗ ui,j−1
 
+ ∆ 2 i,j 2 i,j (5.63)
¯ i,j = 0,
−2(1 + ∆)u

¯ = (∆x/∆y)2 .
where ∆

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 188 / 454
Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

We can solve (5.63) using any of the iterative methods discussed except SOR
(generally need under-relaxation for non-linear problems). However, is (5.63)
diagonally dominant? To be diagonally dominant we must have
¯ − q + ∆
¯ + q ≤ 2(1 + ∆)
¯ ,

|1 − p| + |1 + p| + ∆

where
1 1 ¯ 1/2 vi,j
p= Re ∆x u∗i,j , q= Re ∆x ∆ ∗
.
2 2
¯ then this requires that
Suppose, for example, that p > 1 and q > ∆,
¯ + (∆
(p − 1) + (1 + p) + (q − ∆) ¯ + q) ≤ 2(1 + ∆)
¯

¯
2(p + q) ≤ 2(1 + ∆),

but with p > 1 and q > ∆¯ this condition cannot be satisfied, and equation (5.63)
is not diagonally dominant. The same result holds for p < −1 and q < −∆. ¯
Therefore, we must have |p| ≤ 1 and |q| ≤ ∆¯ or

1 1
Re ∆x u∗i,j ≤ 1, and Re ∆x vi,j ∗ ¯ 1/2

2 2 ≤∆ ,

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 189 / 454

Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

which is a restriction on mesh size for given Re and velocity field.


Three difficulties:
1 As the Reynolds number Re increases, the grid sizes ∆x and ∆y must
decrease.
2 The velocities u∗i,j and vi,j

vary throughout the domain and are unknown.
3 The central difference approximations for the first-order derivatives contribute
to the off-diagonal terms but not the main diagonal terms thereby adversely
affecting diagonal dominance.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 190 / 454
Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Upwind-Downwind Differencing
In order to restore diagonal dominance, we use forward or backward differences for
the first-derivative terms depending upon the signs of the coefficients of the first
derivative terms, i.e. the velocities. For example, consider the u∗ ∂u/∂x term:
1 If u∗i,j > 0, then using a backward difference

∂u ui,j − ui−1,j
u∗ = u∗i,j + O(∆x),
∂x ∆x
which gives a positive addition to the ui,j term to promote diagonal
dominance.
2 If u∗i,j < 0, then using a forward difference

∂u ui+1,j − ui,j
u∗ = u∗i,j + O(∆x),
∂x ∆x
which again gives a positive addition to the ui,j term to promote diagonal
dominance.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 191 / 454

Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Similarly for the v ∗ ∂u/∂y term:



1 If vi,j > 0, then using a backward difference
∂u ∗ ui,j − ui,j−1
v∗ = vi,j + O(∆y).
∂y ∆y

2 If vi,j < 0, then using a forward difference
∂u ∗ ui,j+1 − ui,j
v∗ = vi,j + O(∆y).
∂y ∆y
Let us consider diagonal dominance of the approximations to the x-derivative
terms in Burger’s equation (5.61) in the x-direction
u − ui−1,j , u∗i,j > 0


∂2u ∂u ui+1,j − 2ui,j + ui−1,j Re ui,j  i,j
Tx = −Re u = − ;
∂x2 ∂x (∆x)2 ∆x  ∗
ui+1,j − ui,j , ui,j < 0
therefore,
 ui+1,j + (1 + Re ∆x u∗i,j )ui−1,j − (2 + Re ∆x u∗i,j )ui,j , u∗i,j > 0

(∆x)2 Tx = .
(1 − Re ∆x u∗i,j )ui+1,j + ui−1,j − (2 − Re ∆x u∗i,j )ui,j , u∗i,j <0

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 192 / 454
Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Therefore, if using ADI with splitting of x- and y-derivative terms, diagonal


dominance of the tridiagonal problems along lines of constant y requires that:
1 For u∗i,j > 0 (Re ∆x u∗i,j > 0):

|1| + |1 + Re ∆x u∗i,j | ≤ | − (2 + Re ∆x u∗i,j )|

1 + 1 + Re ∆x u∗i,j = 2 + Re ∆x u∗i,j ,

which is weakly diagonally dominant.


2 For u∗i,j < 0 (Re ∆x u∗i,j < 0):

|1| + |1 − Re ∆x u∗i,j | ≤ | − (2 − Re ∆x u∗i,j )|

1 + 1 − Re ∆x u∗i,j = 2 − Re ∆x u∗i,j ,

which is also weakly diagonally dominant.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 193 / 454

Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Notes:
1 The same is true for the y-derivative terms when sweeping along lines of
constant x.
2 Updwind-downwind differencing forces diagonal dominance; therefore, the
iteration will always converge with no mesh restrictions.
3 The forward and backward differences used for the first-order derivatives are
only first-order accurate, i.e. the method is O(∆x, ∆y) accurate. To see the
potential affects of this error, consider the 1-D Burger’s equation

du d2 u
Re u = 2. (5.64)
dx dx
Recall from section 3.2 that, for example, the first-order, backward-difference
approximation to the first-order derivative is

∆x d2 u
   
du ui − ui−1
= + + ...,
dx i ∆x 2 dx2 i

where we have included the truncation error.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 194 / 454
Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

Substituting into (5.64) gives

∆x d2 u
     2 
∗ ui − ui−1 d u
Re ui + + . . . = ,
∆x 2 dx2 i dx2 i
or   2 
ui − ui−1 Re d u
Re ui = 1− ∆x u∗i .
∆x 2 dx2 i
Therefore, depending upon the values of Re, ∆x, and u∗ , the truncation
error from the first-derivative terms, which is not included in the numerical
solution, may be of the same order, or even larger than, the physical diffusion
term. This is often referred to as artificial or numerical diffusion, the effects
of which increase with increasing Reynolds number.
4 Remedies:
i) Can return to O(∆x2 , ∆y 2 ) accuracy using deferred correction in which we
use the approximate solution to evaluate the leading term of the truncation
error which is then added to the original discretized equation as a source term.
ii) Alternatively, we could use second-order accurate forward and backward
differences, but the resulting system of equations would no longer be
tridiagonal.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 195 / 454

Numerical Solutions of Elliptic Problems Treatment of Nonlinear Convective Terms

5 Note that we have linearized the difference equation, not the differential
equation, in order to obtain a linear system of algebraic equations
⇒ The physical nonlinearity is still being solved for.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 196 / 454
Numerical Solutions of Parabolic Problems Introduction

Outline

6 Numerical Solutions of Parabolic Problems


Introduction
Explicit Methods
Euler Method (First-Order Explicit)
Richardson Method
DuFort-Frankel Method
Stability Analysis
Introduction
Numerical Stability Analysis
Implicit Methods
First-Order Implicit
Crank-Nicolson
Non-Linear Convective Problems
First-Order Explicit
Crank-Nicolson
Upwind-Downwind Differencing
Multidimensional Problems
First-Order Explicit Method
First-Order Implicit Method
ADI Method with Time Splitting

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 197 / 454

Numerical Solutions of Parabolic Problems Introduction

Outline (cont’d)
Factored ADI Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 198 / 454
Numerical Solutions of Parabolic Problems Introduction

Parabolic ⇒ preferred direction of propagation of solution:


1 time (unsteady problems)
2 spatial (some steady problems), e.g. steady duct flows and boundary-layer
flows.
Properties:
1 Initial-value problems, e.g. 1-D, unsteady, heat conduction (cf. elliptic ⇒
BVP)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 199 / 454

Numerical Solutions of Parabolic Problems Introduction

2 Numerical algorithm “marches” in preferred direction in step-by-step manner,


i.e. it is not necessary to keep solution at every time step (just previous
step(s)).
Consider the general linear, 1-D, unsteady equation

∂φ ∂2φ ∂φ
= a(x, t) 2 + b(x, t) + c(x, t)φ + d(x, t). (6.1)
∂t ∂x ∂x
A simple model problem for this is the unsteady, 1-D diffusion equation

∂φ ∂2φ
= α 2, (6.2)
∂t ∂x
where
φ = T ⇒ heat conduction
φ = u ⇒ momentum diffusion (due to viscosity)
φ = ω ⇒ vorticity diffusion
φ = c ⇒ mass diffusion (c = concentration)
Techniques developed for equation (6.2) can be used for equation (6.1).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 200 / 454
Numerical Solutions of Parabolic Problems Introduction

Methods of Solution:
1 Reduce partial differential equation to a set of ordinary differential equations
and solve, e.g. method of lines, predictor-corrector, Runge-Kutta, etc...
2 Finite-difference methods:
a) Explicit methods – obtain equation for φ at each mesh point.
b) Implicit methods – obtain set of algebraic equations for φ at all mesh points at
each ∆t.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 201 / 454

Numerical Solutions of Parabolic Problems Explicit Methods

Outline

6 Numerical Solutions of Parabolic Problems


Introduction
Explicit Methods
Euler Method (First-Order Explicit)
Richardson Method
DuFort-Frankel Method
Stability Analysis
Introduction
Numerical Stability Analysis
Implicit Methods
First-Order Implicit
Crank-Nicolson
Non-Linear Convective Problems
First-Order Explicit
Crank-Nicolson
Upwind-Downwind Differencing
Multidimensional Problems
First-Order Explicit Method
First-Order Implicit Method
ADI Method with Time Splitting

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 202 / 454
Numerical Solutions of Parabolic Problems Explicit Methods

Outline (cont’d)
Factored ADI Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 203 / 454

Numerical Solutions of Parabolic Problems Explicit Methods

Explicit Methods ⇒ Spatial derivatives are all evaluated at previous time level(s),
i.e. single unknown φn+1
i on left-hand-side.
Note that now the superscript n denotes the time step rather than the iteration
number.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 204 / 454
Numerical Solutions of Parabolic Problems Explicit Methods

Euler Method (First-Order Explicit)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 205 / 454

Numerical Solutions of Parabolic Problems Explicit Methods

First-order, forward difference for time derivative

∂φ φn+1 − φni
= i + O(∆t).
∂t ∆t
Second-order, central difference for spatial derivatives at nth time level (known)

∂2φ φni+1 − 2φni + φni−1


2
= 2
+ O(∆x2 ).
∂x (∆x)

Substituting into (6.2)

φn+1
i − φni φni+1 − 2φni + φni−1
=α ,
∆t (∆x)2

and solving for φn+1


i

α∆t
φn+1 = φni + n n n

φ − 2φ + φ
i
(∆x)2 i+1 i i−1
(6.3)
φn+1

i = (1 − 2s)φni + s φni+1 + φni−1 , i = 2, . . . , I,

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 206 / 454
Numerical Solutions of Parabolic Problems Explicit Methods

where s = α∆t/(∆x)2 .
Notes:
1 Equation (6.3) is an explicit equation for φn+1
i at the (n + 1)st time step.
2 Method is second-order accurate in space and first-order accurate in time.
3 Time steps ∆t may be varied from step-to-step.
4 Restrictions on ∆t and ∆x for Euler method applied to 1-D diffusion
equation to remain stable (see section 6.3):
α∆t 1
s= 2
≤ ⇒ stable (very restrictive).
(∆x) 2
α∆t 1
s= 2
> ⇒ unstable.
(∆x) 2

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 207 / 454

Numerical Solutions of Parabolic Problems Explicit Methods

Richardson Method
We want to improve on the temporal accuracy of the Euler method; therefore, we
use a central difference for the time derivative.

φn+1 − φn−1 φn − 2φni + φni−1


i i
= α i+1
2∆t (∆x)2
φn+1 = φn−1 n n n

i i + 2s φ i+1 − 2φ i + φ i−1 , i = 2, . . . , I. (6.4)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 208 / 454
Numerical Solutions of Parabolic Problems Explicit Methods

Notes:
1 Second-order accurate in space and time.
2 Must keep ∆t constant and requires starting method (need φi at two time
steps).
3 Unconditionally unstable for s > 0 ⇒ Do not use.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 209 / 454

Numerical Solutions of Parabolic Problems Explicit Methods

DuFort-Frankel Method
In order to maintain second-order accuracy, but improve stability, let us modify the
Richardson method by taking an average between time levels for φni . To devise
such an approximation, consider the Taylor series approximation at tn+1 about tn
 n n
∆t2 ∂ 2 φ

n+1 n ∂φ
φi = φi + ∆t + + ··· .
∂t i 2 ∂t2 i

Similarly, consider the Taylor series approximation at tn−1 about tn


 n n
∆t2 ∂ 2 φ

n−1 n ∂φ
φi = φi − ∆t + + ··· .
∂t i 2 ∂t2 i

Adding these Taylor series together gives


n
∂2φ

φn+1
i + φn−1
i = 2φni + ∆t 2
+ ··· ,
∂t2 i

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 210 / 454
Numerical Solutions of Parabolic Problems Explicit Methods

and solving for φni leads to


n
∂2φ

1 n+1  1 2
φni = φi + φn−1
i − ∆t + ··· . (6.5)
2 2 ∂t2 i

Therefore, averaging between time levels in this manner is O(∆t2 ) accurate.


Substituting into the difference equation (6.4) from Richardson’s method gives

φn+1 n−1 n+1 n−1


 n  n

i = φ i + 2s φ i+1 − φ i + φ i + φ i−1

(1 + 2s)φn+1 = (1 − 2s)φn−1 + 2s φni+1 + φni−1



i i

1 − 2s n−1 2s
φn+1 φni+1 + φni−1 ,

i = φi + i = 2, . . . , I. (6.6)
1 + 2s 1 + 2s

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 211 / 454

Numerical Solutions of Parabolic Problems Explicit Methods

However, let us consider the consistency of this approximation. Including the


truncation errors of each approximation, the 1-D, unsteady diffusion equation is
approximated as in the Richardson method (with central differences for the
temporal and spacial derivatives)
n
∂2φ φn+1 − φn−1 ∆t2 ∂ 3 φ

∂φ i i
−α 2 = − + ···
∂t ∂x 2∆t 6 ∂t3 i
n
φn − 2φni + φni−1 α ∆x2 ∂4φ

−α i+1 + + ··· .
∆x2 12 ∂x4 i

Substituting the time-averaging equation (6.5) for φni to implement the


DuFort-Frankel method leads to
∂φ ∂2φ φn+1
i − φn−1
i α  n n n+1 n−1

−α 2 = − φ + φ − φ + φ
∂t ∂x 2∆t ∆x2 i+1 i−1 i i

n n n
α ∆t2 ∂ 2 φ ∆t2 ∂ 3 φ α ∆x2 ∂ 4 φ
  
− − + + ··· .
∆x2 ∂t2 i 6 ∂t3 i 12 ∂x4 i

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 212 / 454
Numerical Solutions of Parabolic Problems Explicit Methods

For consistency, all of the truncation error terms must go to zero as


∆t → 0, ∆x → 0. That is, the difference equation must reduce to the differential
equation as ∆x, ∆t → 0. The second and third truncation error terms do so;
however, the first term requires that ∆t → 0 faster than ∆x → 0, i.e. ∆t << ∆x,
for consistency. Because this is not the case in general, the DuFort-Frankel
method is considered inconsistent.
Notes:
1 Second-order accurate in space.
2 Must keep ∆t constant and starting method necessary.
3 Method is unconditionally stable for any s = α∆t/(∆x)2 .
4 The method is inconsistent ⇒ Do not use.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 213 / 454

Numerical Solutions of Parabolic Problems Stability Analysis

Outline

6 Numerical Solutions of Parabolic Problems


Introduction
Explicit Methods
Euler Method (First-Order Explicit)
Richardson Method
DuFort-Frankel Method
Stability Analysis
Introduction
Numerical Stability Analysis
Implicit Methods
First-Order Implicit
Crank-Nicolson
Non-Linear Convective Problems
First-Order Explicit
Crank-Nicolson
Upwind-Downwind Differencing
Multidimensional Problems
First-Order Explicit Method
First-Order Implicit Method
ADI Method with Time Splitting

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 214 / 454
Numerical Solutions of Parabolic Problems Stability Analysis

Outline (cont’d)
Factored ADI Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 215 / 454

Numerical Solutions of Parabolic Problems Stability Analysis

Introduction
Real flows ⇒ small disturbances, e.g. imperfections, vibrations, etc...
Numerical solutions ⇒ small errors, e.g. truncation, round-off, etc...
Issue → What happens to small disturbances/errors as flow and/or solution
evolves in time?
Decay ⇒ stable (disturbances/errors are damped out).
Grow ⇒ unstable (disturbances/errors are amplified).
Two possible sources of instability in CFD:
1 Hydrodynamic instability – the flow itself is inherently unstable (see, for
example, Drazin & Reid)
→ This is real, i.e. physical
2 Numerical instability – the numerical algorithm magnifies small errors
→ This is not physical ⇒ Need a new method.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 216 / 454
Numerical Solutions of Parabolic Problems Stability Analysis

Difficulty → In CFD both are manifest in similar ways, i.e. oscillatory solutions;
therefore, it is often difficult to determine whether oscillatory numerical solutions
are a result of a numerical or hydrodynamic instability.
→ For an example of this, see “Supersonic Boundary-Layer Flow Over a
Compression Ramp,” Cassel, Ruban & Walker, JFM 1995, 1996.
Hydrodynamic vs. Numerical Instability:
1 Hydrodynamic stability analysis (Section 9 and MMAE 514):
Often difficult to perform.
Assumptions must often be made, e.g. parallel flow.
Can provide conclusive evidence for hydrodynamic instability (particularly if
confirmed by analytical or numerical results). For example, in supersonic flow
over a ramp, the Rayleigh Fjørtoft’s theorems are necessary conditions that
can be tested for.
2 Numerical stability analysis:
Often gives guidance, but not always conclusive for complex problems.
Note: Just because a numerical solution does not become oscillatory does not
mean that no physical instabilities are present! There may not be sufficient
resolution.
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 217 / 454

Numerical Solutions of Parabolic Problems Stability Analysis

Numerical Stability Analysis


Two methods:
1) Matrix Method:
More rigorous → evaluates stability of entire scheme, including treatment of
boundary conditions.
More difficult → involves determining eigenvalues of a large matrix.
2) von Neumann (Fourier Series) Method:
For linear initial-value problems with constant coefficients.
⇒ More restrictive, i.e. nonlinear problems must be linearized.
Does not account for boundary conditions.
Most commonly used.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 218 / 454
Numerical Solutions of Parabolic Problems Stability Analysis

Matrix Method
Denote the exact solution of the difference equation at t = tn by φ̃ni ; then the
error is
eni = φni − φ̃ni , i = 2, . . . , I, (6.7)
where φni is the approximate solution at t = tn . Consider the
first-order explicit (Euler) method given by equation (6.3), which is repeated here

φn+1
i = (1 − 2s)φni + s(φni+1 + φni−1 ), (6.8)

where s = α∆t/∆x2 . Both φni and φ̃ni satisfy this equation; thus, the error
satisfies the same equation

ein+1 = (1 − 2s)eni + s(eni+1 + eni−1 ), i = 2, . . . , I. (6.9)

This may be written in matrix form as

en+1 = Aen , n = 0, 1, 2, . . . . (6.10)

Thus, we perform a matrix multiply to advance each time step (cf. matrix form for
iterative methods).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 219 / 454

Numerical Solutions of Parabolic Problems Stability Analysis

The (I − 1) × (I − 1) matrix A and the vector en are

en2
   
1 − 2s s 0 ··· 0 0
 s
 1 − 2s s ··· 0 
 0 


 en3
 0 s 1 − 2s ··· 0  0   en4
n
A= . , e = .
   
.. .. .. .. ..
 .. . . .  .   .
   n 
 0 0 0 ··· 1 − 2s s  eI−1 
0 0 0 ··· s 1 − 2s enI

Note that if φ is specified at the boundaries, then the error is zero there, i.e.
en1 = enI+1 = 0.
The method is stable if the eigenvalues λj of the matrix A are such that

|λj | ≤ 1 for all λj , (6.11)

i.e. ρ ≤ 1, in which case the error will not grow.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 220 / 454
Numerical Solutions of Parabolic Problems Stability Analysis

Because A is tridiagonal with constant elements along each diagonal, the


eigenvalues are (see section 3.5)
 
2 jπ
λj = 1 − (4s) sin , j = 1, . . . , I − 1. (6.12)
2I
Note that there are (I − 1) eigenvalues of the (I − 1) × (I − 1) matrix A. For
equation (6.11) to hold,
 

−1 ≤ 1 − (4s) sin2 ≤ 1.
2I
The right inequality is true for all j (s > 0 and sin2 () > 0). The left inequality is
true if  
2 jπ
1 − (4s) sin ≥ −1
2I
 
2 jπ
−(4s) sin ≥ −2
2I
 
2 jπ 1
s sin ≤
2I 2
   
1 2 jπ
∴ true for all j if s ≤ . 0 < sin <1
2 2I
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 221 / 454

Numerical Solutions of Parabolic Problems Stability Analysis

Thus, the Euler method is stable for s ≤ 1/2.


Notes:
Whereas here we only needed eigenvalues for a tridiagonal matrix, in general
we need to find the eigenvalues of a (I + 1) × (I + 1) matrix.
The effect of different boundary conditions are reflected in A and, therefore,
the resulting eigenvalues.
This is the same method used to obtain convergence properties of iterative
methods for elliptic problems. Recall that the spectral radius ρ(I, J) is the
modulus of the largest eigenvalue of the iteration matrix

φn+1 = Aφn ,

where here n indicates successive iterates rather than time steps. If ρ ≤ 1,


then the iterative procedure converges. In iterative methods, however, we are
concerned not only with whether they will converge or not, but the rate at
which they converge, e.g. Gauss-Seidel vs. Jacobi.
⇒ Minimize ρ for maximum convergence rate.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 222 / 454
Numerical Solutions of Parabolic Problems Stability Analysis

Parabolic problems ⇒ We only require eigenvalues to be less than or equal to


one for stability (it is not necessary to minimize).
⇒ It is often advantageous to use a time-marching (parabolic) scheme for solving
steady (elliptic) problems due to the less restrictive stability criterion. This is
sometimes referred to as the pseudo-transient method, e.g.
∂φ
∇2 φ = 0 → + ∇2 φ = 0.
∂t

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 223 / 454

Numerical Solutions of Parabolic Problems Stability Analysis

von Neumann Method (Fourier Analysis)


We expand the errors along grid lines at one time level as a Fourier series. Then
we determine if the Fourier modes decay or amplify in time.
Expanding the error at t = 0 (n = 0)
I−1
X I−1
X
e(x, 0) = em (x, 0) = am (0)eiθm x , (6.13)
m=1 m=1

where
√ am (0) are the amplitudes of the Fourier modes, θm = mπ, and here
i = −1. At a later time t
I−1
X I−1
X
e(x, t) = em (x, t) = am (t)eiθm x . (6.14)
m=1 m=1

We want to determine how am (t) behaves with time.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 224 / 454
Numerical Solutions of Parabolic Problems Stability Analysis

To do this, define the gain of mode m, Gm (x, t), as

em (x, t) am (t)
Gm (x, t) = = ,
em (x, t − ∆t) am (t − ∆t)

which is the amplification factor for the mth mode during one time step.
Therefore, the error will not grow if |Gm | ≤ 1 for all m, i.e. the method is stable.
If it takes n time steps to get to time t, then the amplification after n time steps is

am (t) am (t − ∆t) am (∆t) am (t)


(Gm )n = ... = ,
am (t − ∆t) am (t − 2∆t) am (0) am (0)

where (Gm )n is the nth power of Gm .


The von Neumann method consists of seeking a solution of the form (6.14), with
am (t) = (Gm )n am (0), of the error equation corresponding to the difference
equation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 225 / 454

Numerical Solutions of Parabolic Problems Stability Analysis

For the first-order explicit (Euler) method, the error equation (4.8) is (use index j
instead of i)

en+1
j = (1 − 2s)enj + s(enj+1 + enj−1 ), j = 2, . . . , I. (6.15)

This equation is linear; therefore, each mode m must satisfy the equation
independently. Thus, substituting (6.14) with am (t) = (Gm )n am (0) into equation
(6.15) gives (canceling am (0) in each term)
h i
n+1 iθm x n iθm x n iθm (x+∆x) iθm (x−∆x)
(Gm ) e = (1 − 2s)(Gm ) e + s(Gm ) e +e .

Dividing by (Gm )n eiθm x we have

(1 − 2s) + s eiθm ∆x + e−iθm ∆x



Gm =
= 1 − 2s [1 − cos(θm ∆x)] [cos(ax) = (eiax + e−iax )/2]
 
2 θm ∆x
Gm = 1 − (4s) sin [sin2 x = [1 − cos(2x)]/2]
2

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 226 / 454
Numerical Solutions of Parabolic Problems Stability Analysis

For stability, |Gm | ≤ 1; therefore,


 
θm ∆x
−1 ≤ 1 − (4s) sin2 ≤1 for all θm = mπ.
2

The right inequality holds for all m and s > 0, i.e.


 
2 θm ∆x
0 ≤ sin ≤ 1.
2

The left inequality holds if s ≤ 1/2 (see matrix method). Thus, for this case we
obtain the same stability criterion as from the matrix method.
Note that |Gm | is the modulus; thus, if we have a complex number, |Gm | equals
the square root of the sum of the squares of the real and imaginary parts.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 227 / 454

Numerical Solutions of Parabolic Problems Implicit Methods

Outline

6 Numerical Solutions of Parabolic Problems


Introduction
Explicit Methods
Euler Method (First-Order Explicit)
Richardson Method
DuFort-Frankel Method
Stability Analysis
Introduction
Numerical Stability Analysis
Implicit Methods
First-Order Implicit
Crank-Nicolson
Non-Linear Convective Problems
First-Order Explicit
Crank-Nicolson
Upwind-Downwind Differencing
Multidimensional Problems
First-Order Explicit Method
First-Order Implicit Method
ADI Method with Time Splitting

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 228 / 454
Numerical Solutions of Parabolic Problems Implicit Methods

Outline (cont’d)
Factored ADI Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 229 / 454

Numerical Solutions of Parabolic Problems Implicit Methods

We can improve on the stability properties of explicit methods by solving implicitly


for more information at the new time level.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 230 / 454
Numerical Solutions of Parabolic Problems Implicit Methods

First-Order Implicit
Recall the first-order explicit method:
Central difference for spatial derivatives on the nth (previous) time level.
First-order forward difference for time derivative.

First-order implicit:
Central difference for spatial derivatives on the (n + 1)st (current) time level.
First-order backward difference for time derivative.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 231 / 454

Numerical Solutions of Parabolic Problems Implicit Methods

For the unsteady, 1-D diffusion equation, this gives

φn+1
j − φnj φn+1 n+1
j+1 − 2φj + φn+1
j−1
=α 2
+ O(∆t, ∆x2 ).
∆t ∆x
Thus,
sφn+1 n+1
j+1 − (1 + 2s)φj + sφn+1 n
j−1 = −φj , j = 2, . . . , I, (6.16)
which is a tridiagonal problem for φn+1
j at the current time level. Note that it is
strongly diagonally dominant.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 232 / 454
Numerical Solutions of Parabolic Problems Implicit Methods

von Neumann Stability Analysis:


Error satisfies equation (6.16)

sen+1 n+1
j+1 − (1 + 2s)ej + sen+1 n
j−1 = −ej , (6.17)

We expand the error at time t as


I−1
X
e(x, t) = (Gm )n am (0)eiθm x , θm = mπ. (6.18)
m=1

Substituting into (6.17) gives (canceling am (0) in each term)

s(Gm )n+1 eiθm (x+∆x) − (1 + 2s)(Gm )n+1 eiθm x


+s(Gm )n+1 eiθm (x−∆x) = −(Gm )n eiθm x
s eiθm ∆x + e−iθm ∆x − (1 + 2s) Gm
  
= −1
[(2s) cos(θm ∆x) − (1 + 2s)] Gm = −1
{1 + 2s [1 − cos(θm ∆x)]} Gm = 1

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 233 / 454

Numerical Solutions of Parabolic Problems Implicit Methods

  −1
θm ∆x
∴ Gm = 1 + (4s) sin2 .
2
Thus, the method is stable if |Gm | ≤ 1. Note that
 
2 θm ∆x
1 + (4s) sin >1
2

for all θm and s > 0. Therefore, Gm < 1, and the method is


unconditionally stable.
Notes:
1) The first-order implicit method is only O(∆t) accurate.
2) There are more computations required per time step than for explicit
methods. However, one can typically use larger time steps (only limited by
temporal resolution required, not stability).
⇒ For explicit methods, we must choose the time step ∆t for both accuracy and
stability, whereas for implicit methods, we must only be concerned with
accuracy.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 234 / 454
Numerical Solutions of Parabolic Problems Implicit Methods

Crank-Nicolson
We prefer second-order accuracy in time; therefore, consider approximating the
equation midway between time levels.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 235 / 454

Numerical Solutions of Parabolic Problems Implicit Methods

For the unsteady, 1-D diffusion equation, this is accomplished as follows

φn+1 − φni α ∂ 2 φn+1 ∂ 2 φn


 
2
i
+ O(∆t ) = 2
+ 2
+ O(∆t2 )
∆t 2 ∂x ∂x
" #
n+1 n+1 n+1
φn+1 − φ n
α φ i+1 − 2φ i + φ i−1 φ n
− 2φ n
+ φ n
i i
= 2
+ i+1 i
2
i−1
.
∆t 2 ∆x ∆x

Later we will show that averaging the diffusion terms across time levels in this
manner is second-order accurate in time. Writing the difference equation in
tridiagonal form, we have

sφn+1 n+1
i+1 − 2(1 + s)φi + sφn+1 n n n
i−1 = −sφi+1 − 2(1 − s)φi − sφi−1 , i = 2, . . . , I,
(6.19)
which we solve for φn+1
i at the current time level.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 236 / 454
Numerical Solutions of Parabolic Problems Implicit Methods

Notes:
1) Second-order accurate in space and time.
2) Unconditionally stable for all s.
3) Apply derivative boundary conditions at current time level.
4) Very popular scheme for parabolic problems.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 237 / 454

Numerical Solutions of Parabolic Problems Implicit Methods

Averaging Across Time Levels:


Here we will again show that averaging across time levels as in the Crank-Nicolson
method is second-order accurate in time. In doing so, we will illustrate a method
to determine the accuracy of a specified approximation (cf. Taylor series
expansions in section 4).
Consider averaging a quantity φi midway between time levels as follows

n+1/2 1 n+1
φi = (φ + φni ) + T.E. (6.20)
2 i

n+1/2
We seek an expression of the form φi = φ̃ + T.E.
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 238 / 454
Numerical Solutions of Parabolic Problems Implicit Methods

Let us expand each term as a Taylor series about (xi , tn+1/2 ):


∞  k
X 1 ∆t
φn+1
i = Dt φ̃,
k! 2
k=0
∞ k ∞ k
(−1)k ∆t
 
X 1 ∆t X
φni = − Dt φ̃ = Dt φ̃,
k! 2 k! 2
k=0 k=0

n+1/2
where Dt = ∂/∂t, and φ̃ is the exact value of φi midway between time levels,
n+1/2
i.e. φi = φ̃ + T.E.. Substituting these expansions into (6.20) gives
∞  k
n+1/2 1X 1  ∆t
1 + (−1)k

φi = Dt φ̃.
2 k! 2
k=0

Note that (
0, k = 1, 3, 5, . . .
1 + (−1)k = .
2, k = 0, 2, 4, . . .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 239 / 454

Numerical Solutions of Parabolic Problems Implicit Methods

Let k = 2m; thus,


∞  2m
n+1/2
X1 ∆t
φi = Dt φ̃
m=0
(2m)! 2
 2  4 .
n+1/2 1 ∆t 1 ∆t
φi = φ̃ + Dt φ̃ + Dt φ̃ + . . .
2! 2 4! 2
The first term in the expansion (m = 0) is the exact value and the second term
(m = 1) gives the truncation error of the approximation. This truncation error
term is
1 2 ∂ 2 φ̃
∆t ;
8 ∂t2
therefore, the approximation (6.20) is second-order accurate, i.e.

n+1/2 1 n+1
φi = (φ + φni ) + O(∆t2 ).
2 i
This shows that averaging across time levels gives an O(∆t2 ) approximation of φi
at the mid-time level tn+1/2 . Note that this agrees with the result (6.5) except for
the constant factor (we averaged across two time levels for the DuFort-Frankel
method).
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 240 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Outline

6 Numerical Solutions of Parabolic Problems


Introduction
Explicit Methods
Euler Method (First-Order Explicit)
Richardson Method
DuFort-Frankel Method
Stability Analysis
Introduction
Numerical Stability Analysis
Implicit Methods
First-Order Implicit
Crank-Nicolson
Non-Linear Convective Problems
First-Order Explicit
Crank-Nicolson
Upwind-Downwind Differencing
Multidimensional Problems
First-Order Explicit Method
First-Order Implicit Method
ADI Method with Time Splitting

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 241 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Outline (cont’d)
Factored ADI Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 242 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Consider the unsteady, 1-D Burger’s equation (1-D, unsteady diffusion equation
with convection terms)
∂u ∂2u ∂u
=ν 2 −u , (6.21)
∂t ∂x ∂x
where ν is the viscosity. We want to consider how the nonlinear convection term
is treated in the various schemes.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 243 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

First-Order Explicit
Approximating spatial derivatives at the previous time level, and using a forward
difference in time, we have

un+1 − uni un − 2uni + uni−1 n n


n ui+1 − ui−1
i
= ν i+1 2
− ui + O(∆x2 , ∆t),
∆t ∆x 2∆x
where the uni in the convection term is known from the previous time level.
Writing in explicit form leads to
   
n+1 Ci Ci
ui = s− uni+1 + (1 − 2s)uni + s + uni−1 ,
2 2
ν∆t un ∆t
where s = ∆x 2 and Ci = ∆x = Courant number. For stability, this method
i

requires that 2 ≤ Re∆x ≤ 2/Ci , where Re∆x = uni ∆x/ν = mesh Reynolds
number. This is very restrictive.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 244 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Crank-Nicolson

un+1 − uni ν ∂ 2 un+1 ∂ 2 un 1 n+1/2 ∂un+1 ∂un


   
i
= + − u +
∆t 2 ∂x2 ∂x2 2 ∂x ∂x
ν n+1 n+1 n+1 n n n

= u − 2u + u + ui+1 − 2ui + u i−1
2∆x2 i+1 i i−1
n+1/2
ui
un+1 n+1 n n 2 2

− i+1 − ui−1 + ui+1 − ui−1 + O(∆x , ∆t ),
4∆x
(6.22)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 245 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

where
1 n+1 n+1/2
(ui + uni ) + O(∆t2 ).
ui = (6.23)
2
Thus, this results in the implicit finite difference equation
   
Ci n+1 n+1 Ci
− s− ui+1 + 2(1 + s)ui − s + un+1
i−1
2 2
    (6.24)
Ci n n Ci
= s− ui+1 + 2(1 − s)ui + s + uni−1 ,
2 2
n+1/2
u ∆t n+1/2
where here Ci = i ∆x , but we do not know ui yet, i.e. it is nonlinear.
Therefore, this procedure requires iteration at each time step:
n+1/2
1) Begin with ui = uni , i.e. use ui from previous time step as initial guess
at current time step.
2) Compute update for un+1
i , i = 1, . . . , I + 1, using equation (6.24).
n+1/2
3) Update ui = 21 (uin+1 + uni ).
4) Repeat until un+1
i converges for all i.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 246 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Notes:
1) It typically requires less than ten iterations to converge at each time step; if
more are required, then the time step ∆t is too large.
2) In elliptic problems we use Picard iteration because we only care about the
final converged solution, whereas here we want an accurate solution at each
time step.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 247 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Upwind-Downwind Differencing
Consider the first-order convective terms in the Crank-Nicolson approximation, i.e.
n+1/2
n+1/2 ∂u
u .
∂x
If un+1/2 > 0, then approximate ∂un+1/2 /∂x as follows:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 248 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Then
" #
∂un+1 ∂un

∂u 1
= + + O(∆t2 )
∂x 2 ∂x i−1/2
∂x i+1/2

" # (6.25)
n+1 n+1
∂u 1 ui − ui−1 uni+1 − uni
= + + O(∆x2 , ∆t2 ),
∂x 2 ∆x ∆x

Notes:
1) Although the finite-difference approximation at the current time level appears
to be a backward difference and that at the previous time level appears to be
a forward difference, they are really central differences evaluated at
half-points in the grid.
2) The fact that this approximation is O(∆x2 , ∆t2 ) accurate will be shown at
the end of this section.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 249 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

If un+1/2 < 0, then approximate ∂un+1/2 /∂x as follows:

Now we have
" #
n+1
n

∂u 1 ∂u ∂u
= + + O(∆t2 )
∂x 2 ∂x i+1/2
∂x i−1/2
" # (6.26)
n+1 n+1 n n
∂u 1 ui+1 − ui u − ui−1
= + i + O(∆x2 , ∆t2 ).
∂x 2 ∆x ∆x

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 250 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Use (6.25) and (6.26) in equation (6.22) rather than central differences
1
un+1 − uni = s un+1 n+1 n+1 n n n

i i+1 − 2u i + ui−1 + ui+1 − 2u i + u i−1
2
n+1/2
(
1 un+1
i − un+1 n n
i−1 + ui+1 − ui , ui >0
− Ci n+1/2
.
2 un+1 − un+1 + un − un , u <0
i+1 i i i−1 i

This gives the tridiagonal problem


( )
un+1
i − un+1
i−1
−sun+1 n+1
i+1 + 2(1 + s)ui − sun+1
i−1 + Ci
un+1
i+1 − ui
n+1

n+1/2
( )
n n
ui+1 − ui , ui >0
= suni+1 + 2(1 − s)uni + suni−1 − Ci .
uni − uni−1 , ui
n+1/2
<0
(6.27)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 251 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Notes:
1) Equation (6.27) is diagonally dominant for all s and Ci (note that Ci may be
positive or negative). Be sure to check this for different equations, i.e. for
other than the one-dimensional Burger’s equation.
2) Iteration at each time step may require under-relaxation on ui ; therefore,
k+1/2
uk+1
i = ωui + (1 − ω)uki , k = 0, 1, 2, . . . ,

where 0 < ω < 1.


3) As shown below, this method is (almost) O(∆x2 , ∆t2 ) accurate.
Truncation Error of Crank-Nicolson with Upwind-Downwind Differencing:
We have used second-order accurate central differences for the ∂u/∂t and
∂ 2 u/∂x2 terms. Therefore, consider the un+1/2 ∂un+1/2 ∂x term with un+1/2 < 0
from (6.26)
∂u 1
= (un+1
i+1 − ui
n+1
+ uni − uni−1 ) + T.E. (6.28)
∂x 2∆x
We would like to determine the order of accuracy, i.e. the truncation error T.E.,
of this approximation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 252 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

n+1/2
Here, Dt = ∂/∂t, Dx = ∂/∂x, and ũ is the exact value of ui midway
between time levels.
We seek an expression of the form
∂u ∂ ũ
= + T.E.
∂x ∂x

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 253 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Expanding each term in equation (6.28) as a 2-D Taylor series about (xi , tn+1/2 ):
∞  k
X 1 ∆t
un+1
i+1 = Dt + ∆xDx ũ,
k! 2
k=0
∞  k
X 1 ∆t
un+1
i = Dt ũ,
k! 2
k=0
∞ k ∞ k
(−1)k ∆t
 
X 1 ∆t X
uni = − Dt ũ = Dt ũ,
k! 2 k! 2
k=0 k=0
∞ k ∞ k
(−1)k ∆t
 
X 1 ∆t X
uni−1 = − Dt − ∆xDx ũ = Dt + ∆xDx ũ,
k! 2 k! 2
k=0 k=0

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 254 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Substituting these expansions into (6.28) gives



(  k
∂u 1 X 1  ∆t
1 − (−1)k

= Dt + ∆xDx
∂x 2∆x k! 2
k=0
 k )
∆t
+ −1 + (−1)k
 
Dt ũ
2

" k  k #
∂u 1 X 1 + (−1)k+1 ∆t ∆t
= Dt + ∆xDx − Dt ũ
∂x 2∆x k! 2 2
k=0

Note that (
0, k = 0, 2, 4, . . .
1 + (−1)k+1 = .
2, k = 1, 3, 5, . . .
Therefore, let k = 2l + 1; thus,

" 2l+1  2l+1 #
∂u 1 X 1 ∆t ∆t
= Dt + ∆xDx − Dt ũ.
∂x ∆x (2l + 1)! 2 2
l=0

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 255 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Recall the binomial theorem


k  
X
k k k−m m
(a + b) = a b ,
m=0
m

k

where m are the binomial coefficients
 
k k!
= (0! = 1).
m m!(k − m)!

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 256 / 454
Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

Thus,

"2l+1 
∂u 1 X 1 X 2l + 1  ∆t 2l+1−m
= Dt (∆xDx )m
∂x ∆x (2l + 1)! m=0 m 2
l=0
 2l+1 #
∆t
− Dt ũ,
2

"2l+1 
1 X 1 X 2l + 1  ∆t 2l−m+1
= Dt (∆xDx )m
∆x (2l + 1)! m=1 m 2
l=0
  2l+1  2l+1 #
2l + 1 ∆t ∆t
+ Dt − Dt ũ,
0 2 2
∞ 2l+1  2l−m+1
X 1 X (2l + 1)! ∆t
= Dt (∆x)m−1 (Dx )m ũ,
(2l + 1)! m=1 m!(2l − m + 1)! 2
l=0
∞ 2l+1  2l−m+1
∂u ∂ ũ X X 1 ∆t
= + Dt (∆x)m−1 (Dx )m ũ.
∂x ∂x m=1
m!(2l − m + 1)! 2
l=1

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 257 / 454

Numerical Solutions of Parabolic Problems Non-Linear Convective Problems

To obtain the truncation error, consider the l = 1 term (m = 1, 2, 3)


 2  
1 ∆t 1 ∆t 1
Dt2 Dx ũ + ∆xDt Dx2 ũ + ∆x2 Dx3 ũ.
1!2! 2 2!1! 2 3!1!

Therefore, the truncation error is

O(∆t2 , ∆t∆x, ∆x2 ).

Thus, if ∆t < ∆x then the approximation is O(∆x2 ) accurate, and if ∆t > ∆x


then the approximation is O(∆t2 ) accurate. This is better than O(∆x) or O(∆t),
but strictly speaking it is not O(∆t2 , ∆x2 ).
Similarly, the v∂u/∂y term has truncation error

O(∆t2 , ∆t∆y, ∆y 2 ).

The loss of accuracy is due to the diagonal averaging across time levels.
Note:
1 Method is unconditionally stable.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 258 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

Outline

6 Numerical Solutions of Parabolic Problems


Introduction
Explicit Methods
Euler Method (First-Order Explicit)
Richardson Method
DuFort-Frankel Method
Stability Analysis
Introduction
Numerical Stability Analysis
Implicit Methods
First-Order Implicit
Crank-Nicolson
Non-Linear Convective Problems
First-Order Explicit
Crank-Nicolson
Upwind-Downwind Differencing
Multidimensional Problems
First-Order Explicit Method
First-Order Implicit Method
ADI Method with Time Splitting

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 259 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

Outline (cont’d)
Factored ADI Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 260 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

Consider the unsteady, 2-D diffusion equation


 2
∂ φ ∂2φ

∂φ
=α + 2 , φ = φ(x, y, t), (6.29)
∂t ∂x2 ∂y

with initial condition

φ(x, y, 0) = φ0 (x, y) at t = 0, (6.30)

and boundary conditions (Dirichlet, Neumann or mixed) on a closed contour C


enclosing the domain. We discretize as follows (t is normal to x − y plane):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 261 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

First-Order Explicit Method

Approximating equation (6.29) using a forward difference in time and central


differences in space at the previous time level gives

φn+1 n  n
i,j − φi,j φi+1,j − 2φni,j + φni−1,j φni,j+1 − 2φni,j + φni,j−1

=α +
∆t (∆x)2 (∆y)2
+O(∆t, ∆x2 , ∆y 2 ).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 262 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

Therefore, solving for the only unknown gives the explicit expression

φn+1 n n n n n
i,j = (1 − 2sx − 2sy )φi,j + sx (φi+1,j + φi−1,j ) + sy (φi,j+1 + φi,j−1 ), (6.31)

where sx = α∆t/(∆x)2 and sy = α∆t/(∆y)2 .


For numerical stability, a von Neumann stability analysis requires that
1
sx + sy ≤ .
2
Thus, for example if ∆x = ∆y, i.e. sx = sy = s, we must have

1
s≤ ,
4
1
which is even more restrictive than for the 1-D diffusion equation where s ≤ 2 for
stability.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 263 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

First-Order Implicit Method

Applying a backward difference in time and central differences in space at the


current time level leads to the implicit expression

(1 + 2sx + 2sy )φn+1 n+1 n+1 n+1 n+1 n


i,j − sx (φi+1,j + φi−1,j ) − sy (φi,j+1 + φi,j−1 ) = φi,j . (6.32)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 264 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

Notes:
1 Unconditionally stable for all sx and sy .
2 Crank-Nicolson could be used to obtain second-order accuracy in time. It
produces a similar implicit equation, but with more terms on the
right-hand-side, i.e. evaluated at the previous time step.
3 Produces a banded matrix (with five unknowns) that is difficult to solve
efficiently.

4 Alternatively, we could split each time step into two half steps, called ADI
with time splitting, resulting in two sets of tridiagonal problems per time step:
Step 1: Solve implicitly for terms associated with one coordinate direction.
Step 2: Solve implicitly for terms associated with other coordinate direction.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 265 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

ADI Method with Time Splitting (Fractional-Step Method)


As mentioned above, we split each time step into two half-time steps as follows:
1 Sweep along lines of constant y during the first half step.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 266 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

n+1/2
" n+1/2 n+1/2 n+1/2
#
φi,j − φni,j φi+1,j − 2φi,j + φi−1,j φni,j+1 − 2φni,j + φni,j−1
=α + .
∆t/2 (∆x)2 (∆y)2
(6.33)
Therefore,
1 n+1/2 n+1/2 1 n+1/2 1 1
sx φi+1,j −(1+sx )φi,j + sx φi−1,j = − sy φni,j+1 −(1−sy )φni,j − sy φni,j−1 .
2 2 2 2
(6.34)
n+1/2
The tridiagonal problems (6.34) are solved for φi,j , i = 1, . . . , I + 1,
j = 1, . . . , J + 1, at the intermediate time level.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 267 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

2 Sweep along lines of constant x during the second half step.

n+1/2 n+1/2 n+1/2 n+1/2


" #
φn+1
i,j − φi,j φi+1,j − 2φi,j + φi−1,j φn+1
i,j+1 − 2φ n+1
i,j + φ n+1
i,j−1
=α 2
+ 2
.
∆t/2 (∆x) (∆y)
(6.35)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 268 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

Therefore,
1 n+1 1 1 n+1/2 n+1/2 1 n+1/2
sy φn+1 n+1
i,j+1 −(1+sy )φi,j + sy φi,j−1 = − sx φi+1,j −(1−sx )φi,j − sx φi−1,j .
2 2 2 2
(6.36)
n+1
The tridiagonal problems (6.36) are solved for φi,j , i = 1, . . . , I + 1,
j = 1, . . . , J + 1, at the current time level.
Notes:
1 Method is O(∆t2 , ∆x2 , ∆y 2 ).
2 Requires boundary conditions at the intermediate time level n + 1/2 for
equation (6.34).
For example, if the boundary condition at x = 0 is Dirichlet as follows

φ(0, y, t) = a(y, t),

then
φn1,j = anj .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 269 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

Subtracting equation (6.35) from (6.33) gives


n+1/2 n+1/2
" #
φi,j − φni,j φn+1
i,j − φ i,j φ n
i,j+1 − 2φ n
i,j + φ n
i,j−1 φ n+1
i,j+1 − 2φ n+1
i,j + φ n+1
i,j−1
− =α 2
− 2
∆t/2 ∆t/2 (∆y) (∆y)

or solving for the unknown at the intermediate time level results in

n+1/2 1 n  1  n
φi,j + φn+1 + sy φi,j+1 − 2φni,j + φni,j−1 − φn+1 n+1 n+1

φi,j = i,j i,j+1 − 2φi,j + φi,j−1 .
2 4
Applying this equation at the boundary x = 0, leads to

n+1/2 1 n  1 
aj + ajn+1 + sy anj+1 − 2anj + anj−1 − an+1 n+1 n+1

φ1,j = j+1 − 2aj + aj−1 .
2 4
This provides the boundary condition for φ1,j at the intermediate (n + 1/2)
time level. Note that the first term on the right-hand-side is the average of a
at the n and n + 1 time levels, the second term is ∂ 2 an /∂y 2 , and the third
term is ∂ 2 an+1 /∂y 2 .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 270 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

3 For stability, apply von Neumann analysis at each half step and take the
product of the resulting amplification factors, G1 and G2 , to obtain G for the
full time step.
⇒ Method is unconditionally stable for all sx and sy .
4 In 3-D, we require three fractional steps (∆t/3) for each time step, and the
method is only conditionally stable, where

sx , sy , sz ≤ 1.5,

for stability (sz = α∆t/(∆z)2 ).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 271 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

Factored ADI Method


Let us reconsider the unsteady, 2-D diffusion equation
 2
∂ φ ∂2φ

∂φ
=α + 2 , (6.37)
∂t ∂x2 ∂y

and apply the Crank-Nicolson approximation

φn+1 n
i,j − φi,j α  2 n+1
δx φi,j + δx2 φni,j + δy2 φn+1 2 n

= i,j + δy φi,j ,
∆t 2
where δ represents second-order central difference operators (as in section 5.7)

φi+1,j − 2φi,j + φi−1,j


δx2 φi,j = ,
(∆x)2

φi,j+1 − 2φi,j + φi,j−1


δy2 φi,j = .
(∆y)2

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 272 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

Rewriting the difference equation with the unknowns on the left-hand-side and the
knowns on the right leads to
   
1 1
1 − α∆t δx2 + δy2 φn+1 2 2
φni,j .
 
i,j = 1 + α∆t δx + δy (6.38)
2 2

We “factor” the difference operator on the left-hand-side as follows


    
1 2 2
 1 2 1 2
1 − α∆t δx + δy ≈ 1 − α∆tδx 1 − α∆tδy , (6.39)
2 2 2

where the first factor only involves the difference operator in the x-direction, and
the second factor only involves the difference operator in the y-direction. The
factored operator produces an extra term as compared to the unfactored operator
1 2 2 2 2
α ∆t δx δy = O(∆t2 ),
4
which is O(∆t2 ). Therefore, the factorization (6.39) is consistent with the
second-order accuracy in time of the Crank-Nicolson approximation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 273 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

The factored form of (6.38) is


    
1 1 1
1 − α∆tδx2 1 − α∆tδy2 φn+1 2 2
φni,j ,

i,j = 1 + α∆t δx + δy
2 2 2

which can be solved in two steps by defining the intermediate variable


 
1
φ̂i,j = 1 − α∆tδy φn+1
2
i,j . (6.40)
2
n+1/2
Note that φ̂i,j is not the same as φi,j , which is an intermediate approximation
to φi,j at the half time step.
The two stage solution process is:
1 Sweep along constant y-lines solving
   
1 2 1 2 2
 n
1 − α∆tδx φ̂i,j = 1 + α∆t δx + δy φi,j , (6.41)
2 2

which produces a tridiagonal problem at each j for φ̂i,j , i = 1, . . . , I + 1.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 274 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

2 Sweep along constant x-lines solving (from (6.40))


 
1
1 − α∆tδy2 φn+1 i,j = φ̂i,j , (6.42)
2

which produces a tridiagonal problem at each i for φn+1


i,j , j = 1, . . . , J + 1 at
the current time step. Note that the right-hand-side of this equation is the
solution of (6.41).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 275 / 454

Numerical Solutions of Parabolic Problems Multidimensional Problems

Notes:
1 Similar to the ADI method with time splitting, but have an intermediate
n+1/2
variable φ̂i,j rather than half time step φi,j .
Factored ADI is somewhat faster; it only requires one evaluation of the spatial
derivatives on the right-hand-side per time step (for equation (6.41)) rather
than two for the ADI method (see equations (6.34) and (6.36)).
2 Method is O(∆t2 , ∆x2 , ∆y 2 ) accurate and is unconditionally stable (even for
3-D implementation of unsteady diffusion equation).
3 Requires boundary conditions for the intermediate variable φ̂i,j to solve
(6.41). These are obtained from equation (6.40) applied at the boundaries
(see Fletcher, section 8.4.1).
4 The order of solution can be reversed, i.e. we could define
 
1
φ̂i,j = 1 − α∆tδx φn+1
2
i,j .
2

instead of (6.40).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 276 / 454
Numerical Solutions of Parabolic Problems Multidimensional Problems

5 If we have non-linear convective terms, i.e. unsteady Navier-Stokes or


transport equation, use upwind-downwind differencing as in section 6.5.3. See
Peridier, Smith & Walker, JFM, Vol. 232, pp. 99–131 (1991), which shows
factored ADI with upwind-downwind differencing applied to the the unsteady
boundary-layer equations (but method applies to more general equations).
6 Observe that numerical methods for parabolic problems do not require
relaxation or acceleration parameters as is often the case for elliptic solvers.
Recall that the issue is numerical stability, not iterative convergence.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 277 / 454

Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Outline

7 Numerical Solutions of Navier-Stokes Equations


Primitive-Variables Formulation
Vorticity-Streamfunction Formulation
Boundary Conditions for Vorticity-Streamfunction Formulation
Thom’s Method
Jensen’s Method
Numerical Solutions of Coupled Systems of Equations
Sequential Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 278 / 454
Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Incompressible flow of a Newtonian fluid is governed by the Navier-Stokes


equations, given by the momentum equation

∂V∗
 
ρ ∗
+ V · ∇V = −∇p∗ + µ∇2 V∗ ,
∗ ∗
∂t

and the continuity equation


∇ · V∗ = 0.
The density is ρ, the viscosity is µ, the pressure is p∗ (x∗ , y ∗ , z ∗ , t∗ ), and the
velocity vector in Cartesian coordinates, for example, is V∗ = u∗ i + v ∗ j + w∗ k.
It is convenient to nondimensionalize using a characteristic length scale L and
velocity U according to

(x∗ , y ∗ , z ∗ ) t∗ V∗ p∗
(x, y, z) = , t= V= , p= .
L L/U U ρU 2

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 279 / 454

Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Applying to the Navier-Stokes equations gives


∂V 1 2
+ V · ∇V = −∇p + ∇ V,
∂t Re
and
∇ · V = 0,
where Re = ρU L/µ is the nondimensional Reynolds number.
In 2-D, Cartesian coordinates, the incompressible Navier-Stokes equations are:
x-momentum:
1 ∂2u ∂2u
 
∂u ∂u ∂u ∂p
+u +v =− + + 2 , (7.1)
∂t ∂x ∂y ∂x Re ∂x2 ∂y
y-momentum:

1 ∂2v ∂2v
 
∂v ∂v ∂v ∂p
+u +v =− + + 2 , (7.2)
∂t ∂x ∂y ∂y Re ∂x2 ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 280 / 454
Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Continuity:
∂u ∂v
+ = 0. (7.3)
∂x ∂y
Thus, we have three coupled equations for three dependent variables
u(x, y, t), v(x, y, t) and p(x, y, t), which we refer to as primitive variables.
Therefore, the system is closed mathematically:
Given v(x, y, t) and p(x, y, t), we can determine u(x, y, t) from equation
(7.1).
Given u(x, y, t) and p(x, y, t), we can determine v(x, y, t) from equation
(7.2).
But how do we determine p(x, y, t) given that p does not appear in equation
(7.3)?
⇒ Need equation for p(x, y) in terms of u(x, y) and v(x, y) at time t.
To obtain such an equation, take the divergence of the momentum equation in
vector form, i.e. ∇ · (N S), which in 2-D is equivalent to taking ∂/∂x of equation
(7.1), ∂/∂y of equation (7.2) and adding.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 281 / 454

Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Doing so gives
" 2
2 2
∂2u ∂ 2 u ∂v ∂u ∂2u

∂ p ∂ p ∂u
2
+ 2 = − + +u 2 + +v
∂x ∂y ∂t∂x ∂x ∂x ∂x ∂y ∂x∂y
#
2 2
  2 2
∂ v ∂u ∂v ∂ v ∂v ∂ v
+ + +u + +v 2
∂t∂y ∂y ∂x ∂x∂y ∂y ∂y
1 ∂3u ∂3u ∂3v ∂3v
 
+ + + + .
Re ∂x3 ∂x∂y 2 ∂x2 ∂y ∂y 3

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 282 / 454
Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Doing so gives
" 2
2 2
∂2u ∂ 2 u ∂v ∂u ∂2u

∂ p ∂ p ∂u
2
+ 2 = − + +u 2 + +v
∂x ∂y ∂t∂x ∂x ∂x ∂x ∂y ∂x∂y
#
2 2
  2 2
∂ v ∂u ∂v ∂ v ∂v ∂ v
+ + +u + +v 2
∂t∂y ∂y ∂x ∂x∂y ∂y ∂y
1 ∂3u ∂3u ∂3v ∂3v
 
+ + + + .
Re ∂x3 ∂x∂y 2 ∂x2 ∂y ∂y 3

Or
∂2p ∂2p
      
∂ ∂u ∂v ∂ ∂u ∂v ∂ ∂u ∂v
+ = − + +u + +v +
∂x2 ∂y 2 ∂t ∂x ∂y ∂x ∂x# ∂y ∂y ∂x ∂y
 2  2
∂u ∂v ∂v ∂u
+ + +2
∂x ∂y ∂x ∂y
 2 
∂ 2 ∂u ∂v
  
1 ∂ ∂u ∂v
+ + + 2 + .
Re ∂x2 ∂x ∂y ∂y ∂x ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 283 / 454

Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

But from the continuity equation (7.3)


0 0 0

2 2 *
 *
 *

     
∂ p ∂ p  ∂ ∂u ∂v  ∂ ∂u ∂v  ∂ ∂u ∂v 
+ = − + + u + + v +
∂x2 ∂y 2 ∂t∂x ∂y ∂x∂x ∂y ∂y ∂x ∂y

     

 2  2 #
∂u ∂v ∂v ∂u
+ + +2
∂x ∂y ∂x ∂y
0 0
 
2 *
 2 *

   
1  ∂ ∂u ∂v  ∂ ∂u ∂v 

+  2 + + 2 + .
Re ∂x ∂x ∂y ∂y ∂x  ∂y
 

We also have from continuity that


 2  2
∂u ∂v ∂u ∂v
= =− .
∂x ∂y ∂x ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 284 / 454
Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Substituting gives
∂2p ∂2p
 
∂u ∂v ∂v ∂u
+ 2 =2 − , (7.4)
∂x2 ∂y ∂x ∂y ∂x ∂y
which is a Poisson equation for pressure with u(x, y) and v(x, y) known from
solutions of equations (7.1) and (7.2), respectively.
Notes:
1 The unsteady momentum equations (7.1) and (7.2) are parabolic in time.
⇒ May be solved using Crank-Nicolson, ADI, Factored-ADI, etc....
2 The pressure equation (7.4) and steady forms of (7.1) and (7.2) are elliptic.
⇒ Can be solved using cyclic reduction, Gauss-Seidel, ADI, multigrid, etc....

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 285 / 454

Numerical Solutions of Navier-Stokes Equations Primitive-Variables Formulation

Velocity Boundary Conditions (us and un are velocity components tangential and
normal to the surface, respectively):
Surface: us = un = 0 (no slip and impermeability)
Inflow: us and un specified
Outflow: ∂u ∂un
∂n = ∂n = 0 (fully-developed flow)
s

∂us
Symmetry: ∂n = 0, un = 0 (no flow through symmetry plane)
Note that the domain must be sufficiently long for the fully-developed outflow
boundary condition to be valid.
Pressure Boundary Conditions at a Surface:
From the momentum equations (7.1) and (7.2) with u = v = 0:
∂p 1 ∂2u
n=x⇒ =
∂x Re ∂x2
∂p 1 ∂2v
n=y⇒ =
∂y Re ∂x2
Other boundary conditions for pressure obtained similarly.
Observe that we have Neumann boundary conditions on pressure at solid surfaces.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 286 / 454
Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

Outline

7 Numerical Solutions of Navier-Stokes Equations


Primitive-Variables Formulation
Vorticity-Streamfunction Formulation
Boundary Conditions for Vorticity-Streamfunction Formulation
Thom’s Method
Jensen’s Method
Numerical Solutions of Coupled Systems of Equations
Sequential Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 287 / 454

Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

The streamfunction ψ(x, y, t) is defined for 2-D, incompressible flow by

∂ψ ∂ψ
u= , v=− , (7.5)
∂y ∂x

such that the continuity equation (7.3) is identically satisfied. Lines of constant ψ
are called streamlines and are everywhere tangent to the local velocity vectors.
The vorticity ω(x, y, t) in 2-D is defined by

∂v ∂u
ω= − , (7.6)
∂x ∂y
and measures the local rate of rotation of fluid particles, with the sign
corresponding to the right-hand-rule. Note that in general 3-D flows vorticity is a
vector →

ω = ∇ × V.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 288 / 454
Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

Note that vorticity 6= vortex:


vorticity – spatial distribution of rotation rates of individual fluid particles.
vortex – a spatially coherent region of fluid rotating as a unit, i.e. an “eddy.”
Example of a flow with vorticity, but no vortex?
Example of a flow with vortex, but no vorticity?
To obtain the vorticity-transport equation, take the curl of the momentum
equation in vector form, i.e. ∇ × (N S), which in 2-D is equivalent to taking ∂/∂x
of equation (7.2) minus ∂/∂y of equation (7.1):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 289 / 454

Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

Doing so gives

∂2v ∂u ∂v ∂2v ∂v ∂v ∂2v


+ +u 2 + +v
∂x∂t ∂x ∂x ∂x ∂x ∂y ∂x∂y
2 2
∂ u ∂u ∂u ∂ u ∂v ∂u ∂2u
− − −u − −v 2
∂y∂t ∂y ∂x ∂x∂y ∂y ∂y ∂y
∂2p ∂2p 1 ∂3v ∂3v ∂3u ∂3u
 
=− + + + − − 3 .
∂x∂y ∂x∂y Re ∂x3 ∂x∂y 2 ∂x2 ∂y ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 290 / 454
Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

Doing so gives

∂2v ∂u ∂v ∂2v ∂v ∂v ∂2v


+ +u 2 + +v
∂x∂t ∂x ∂x ∂x ∂x ∂y ∂x∂y
2 2
∂ u ∂u ∂u ∂ u ∂v ∂u ∂2u
− − −u − −v 2
∂y∂t ∂y ∂x ∂x∂y ∂y ∂y ∂y
∂2p ∂2p 1 ∂3v ∂3v ∂3u ∂3u
 
=− + + + − − 3 .
∂x∂y ∂x∂y Re ∂x3 ∂x∂y 2 ∂x2 ∂y ∂y

Or
       
∂ ∂v ∂u ∂ ∂v ∂u ∂ ∂v ∂u ∂v ∂u ∂v
− +u − +v − + +
∂t ∂x ∂y ∂x ∂x ∂y ∂y ∂x ∂y ∂x ∂x ∂y
 2 
∂ 2 ∂v
    
∂u ∂u ∂v 1 ∂ ∂v ∂u ∂u
− + = − + 2 −
∂y ∂x ∂y Re ∂x2 ∂x ∂y ∂y ∂x ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 291 / 454

Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

But from the continuity equation (7.3) and the definition of vorticity (7.6)

 *

 ω  *

 ω  *
 ω  *

 0
∂ ∂v ∂u  ∂ ∂v ∂u  ∂ ∂v ∂u  ∂v ∂u ∂v 
− +u − +v − + +
∂t∂x ∂y ∂x∂x ∂y ∂y ∂x ∂y ∂x∂x ∂y
    
  

0  ω ω
*
 2 *
 2 *

     
∂u ∂u ∂v  1  ∂ ∂v ∂u ∂ ∂v ∂u

− + = − + − 
∂y ∂x ∂y Re ∂x2 ∂x ∂y ∂y 2 ∂x ∂y
  
  

Therefore, the vorticity-transport equation in 2-D is

1 ∂2ω ∂2ω
 
∂ω ∂ω ∂ω
+u +v = + , (7.7)
∂t ∂x ∂y Re ∂x2 ∂y 2

which is a convection-diffusion equation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 292 / 454
Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

The vorticity and streamfunction may be related by substituting (7.5) into (7.6)
to obtain
∂2ψ ∂2ψ
+ = −ω, (7.8)
∂x2 ∂y 2
which is a Poisson equation for ψ(x, y, t) if ω(x, y, t) is known.
Notes:
1 Equations (7.7) and (7.8) are coupled equations for ω and ψ (with
u = ∂ψ/∂y, v = −∂ψ/∂x in (7.7)).
2 The vorticity-transport equation (7.7) is parabolic when unsteady and elliptic
when steady. The streamfunction equation (7.8) is elliptic.
3 The pressure terms have been eliminated, i.e. there is no need to calculate
the pressure in order to advance the solution in time.
→ Can compute p(x, y, t) from equation (7.4) if desired.
4 The vorticity-streamfunction formulation consists of two equations for the
two unknowns ω(x, y, t) and ψ(x, y, t) (cf. primitive variables formulation
with three equations for three unknowns in 2-D).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 293 / 454

Numerical Solutions of Navier-Stokes Equations Vorticity-Streamfunction Formulation

5 Unlike the primitive variables formulation, the ω–ψ formulation does not
easily extend to 3-D:
Three components of vorticity ⇒ Three vorticity equations.
Stretching and tilting terms in 3-D vorticity equations.
Cannot define streamfunction in 3-D ⇒ Vorticity-velocity potential
formulation.
6 We do not have straightforward boundary conditions for vorticity (see next
section).

⇒ Because of notes (3) and (4) (notwithstanding (6)), this is the preferred
formulation for 2-D, incompressible flows.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 294 / 454
Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

Outline

7 Numerical Solutions of Navier-Stokes Equations


Primitive-Variables Formulation
Vorticity-Streamfunction Formulation
Boundary Conditions for Vorticity-Streamfunction Formulation
Thom’s Method
Jensen’s Method
Numerical Solutions of Coupled Systems of Equations
Sequential Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 295 / 454

Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

Consider the driven cavity flow:

Recall that the streamfunction is defined by


∂ψ ∂ψ
u= , v=− .
∂y ∂x

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 296 / 454
Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

Thus, on AD and BC:


∂ψ
u=0⇒ = 0 ⇒ ψ = const. = 0
∂y
∂ψ
v=0⇒ =0
∂x
∂ψ
∴ ψ and specified (n = x).
∂n
Similarly, on AB:
∂ψ
v=0⇒ =0⇒ψ=0
∂x
∂ψ
u=0⇒ =0
∂y
∂ψ
∴ ψ and specified (n = y).
∂n

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 297 / 454

Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

On CD:
∂ψ
v=0⇒ =0⇒ψ=0
∂x
∂ψ
u=1⇒ =1
∂y
∂ψ
∴ ψ and
specified (n = y).
∂n
At solid boundaries, therefore, the streamfunction and its normal derivative are
specified. Note that we have two boundary conditions on the streamfunction,
where only one is needed.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 298 / 454
Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

In order to consider boundary conditions on vorticity, recall its definition


∂v ∂u
ω= − .
∂x ∂y
Thus, for example, on AB:
∂v ∂u τw
v=0⇒ =0⇒ω=− =− ,
∂x ∂y µ
where τw is the wall shear stress. However, τw is unknown and must be
determined from the solution.
⇒ Don’t have boundary condition for ω.
Strategy:
Use ψ = const. as boundary condition for streamfunction equation (7.8).
Use “extra” boundary condition ∂ψ/∂n = const. to obtain a condition for ω
at the surface.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 299 / 454

Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

Thom’s Method
Consider, for example, the lower boundary AB (y = 0). Throughout the domain
we have
∂2ψ ∂2ψ
+ = −ω.
∂x2 ∂y 2
∂2ψ
However, along AB ψ = 0; therefore, = 0, and
∂x2

∂ 2 ψ

ωw = − . (7.9)
∂y 2 y=0

We also have
∂ψ
u= = 0 on y = 0.
∂y
For generality, consider a moving wall with

∂ψ
= uw (x) = g(x), (7.10)
∂y y=0

where the tangential wall velocity uw (x) = g(x) is specified.


MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 300 / 454
Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

Applying a central difference to (7.10) gives

ψi,2 − ψi,0
= gi + O(∆y 2 );
2∆y
therefore,
ψi,0 = ψi,2 − 2∆y gi + O(∆y 3 ). (7.11)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 301 / 454

Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

A central difference for (7.9) gives

ψi,2 − 2ψi,1 + ψi,0


ωi,1 = − + O(∆y 2 ). (7.12)
(∆y)2

Substituting (7.11) for ψi,0 in the previous equation leads to

2
ωi,1 = − [ψi,2 − ψi,1 − ∆y gi ] + O(∆y). (7.13)
(∆y)2

We then use the most recent iterate for streamfunction ψ to obtain a Dirichlet
boundary condition for vorticity ω.
Note: The truncation error for Thom’s method is only O(∆y). However, it
exhibits second-order convergence (Huang & Wetton 1996).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 302 / 454
Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

Jensen’s Method
We would like a method that is O(∆y 2 ) overall; therefore, consider using an
O(∆y 3 ) approximation for (7.10)

−2ψi,0 − 3ψi,1 + 6ψi,2 − ψi,3


= gi + O(∆y 3 );
6∆y
therefore,
3 1
ψi,0 = − ψi,1 + 3ψi,2 − ψi,3 − 3∆y gi + O(∆y 4 ).
2 2
Substituting into (7.12) results in
1
ωi,1 = − 2
[7ψi,1 − 8ψi,2 + ψi,3 + 6∆y gi ] + O(∆y 2 ). (7.14)
2(∆y)

We then use (7.14) in place of (7.13) to obtain a Dirichlet boundary condition for
vorticity.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 303 / 454

Numerical Solutions of Navier-Stokes Equations Boundary Conditions for Vorticity-Streamfunction Formulation

Notes:
1 For discussion of boundary conditions on vorticity and streamfunction at
inlets and outlets, see Fletcher, Vol. II, pp. 380–381.
2 Treatment of vorticity (and pressure) boundary conditions is still an active
area of research and debate (see, for example, Rempfer 2003).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 304 / 454
Numerical Solutions of Navier-Stokes Equations Numerical Solutions of Coupled Systems of Equations

Outline

7 Numerical Solutions of Navier-Stokes Equations


Primitive-Variables Formulation
Vorticity-Streamfunction Formulation
Boundary Conditions for Vorticity-Streamfunction Formulation
Thom’s Method
Jensen’s Method
Numerical Solutions of Coupled Systems of Equations
Sequential Method

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 305 / 454

Numerical Solutions of Navier-Stokes Equations Numerical Solutions of Coupled Systems of Equations

Until now, we have considered methods for solving single equations; but in fluid
dynamics we must solve systems of coupled equations, such as the Navier-Stokes
equations in the primitive-variables or vorticity-streamfunction formulations.
Two methods for treating coupled equations numerically:
1 Sequential Solution:
Solve, i.e. iterate on, each equation for its dominant variable, treating the
other variables as known, i.e. use most recent values.
Requires one pass through mesh for each equation at each iteration.
⇒ Most common and easiest to implement.
2 Simultaneous (or Coupled) Solution:
Combine coupled equations into a single system of algebraic equations.
⇒ If have n dependent variables (e.g. 2-D primitive variables ⇒ n = 3 (u, v, p)),
produces an n × n block tridiagonal system of equations that is solved for all
the dependent variables simultaneously.
See, for example, S. P. Vanka, J. Comput. Phys., Vol. 65, pp. 138–158 (1986).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 306 / 454
Numerical Solutions of Navier-Stokes Equations Numerical Solutions of Coupled Systems of Equations

For example, if solving using Gauss-Seidel:


1 Sequential ⇒ One Gauss-Seidel expression for each dependent variable
updated in succession.
2 Simultaneous ⇒ Solve n equations for n unknown dependent variables
directly at each grid point.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 307 / 454

Numerical Solutions of Navier-Stokes Equations Numerical Solutions of Coupled Systems of Equations

Sequential Method
Consider for example the primitive-variables formulation.
Steady Problems:

Note:
May require underrelaxation for convergence due to nonlinearity.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 308 / 454
Numerical Solutions of Navier-Stokes Equations Numerical Solutions of Coupled Systems of Equations

Unsteady Problems:

Notes:
Outer loop for time marching.
Inner loop to obtain solution of coupled equations at current time step.
⇒ Trade-off: Generally, reducing ∆t reduces number of inner loop iterations. (If
∆t small enough, no iteration is necessary).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 309 / 454

Grids and Grid Generation Introduction

Outline

8 Grids and Grid Generation


Introduction
Staggered vs. Collocated Grids
Uniform vs. Non-Uniform Grids
Grid Generation
Algebraic Grid Generation
Elliptic Grid Generation
Variational Grid Generation

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 310 / 454
Grids and Grid Generation Introduction

Thus far we have used what are called “uniform, collocated grids.”
Uniform ⇒ Grid spacings in each direction ∆x and ∆y are uniform.
Collocated ⇒ All dependent variables are approximated at the same point.
In the following two sections we consider alternatives to uniform and collocated
grids.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 311 / 454

Grids and Grid Generation Staggered vs. Collocated Grids

Outline

8 Grids and Grid Generation


Introduction
Staggered vs. Collocated Grids
Uniform vs. Non-Uniform Grids
Grid Generation
Algebraic Grid Generation
Elliptic Grid Generation
Variational Grid Generation

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 312 / 454
Grids and Grid Generation Staggered vs. Collocated Grids

Recall the 2-D, incompressible Navier-Stokes equations in the primitive variables


formulation
1 ∂2u ∂2u
 
∂u ∂u ∂u ∂p
+u +v =− + + 2 , (8.1)
∂t ∂x ∂y ∂x Re ∂x2 ∂y
1 ∂2v ∂2v
 
∂v ∂v ∂v ∂p
+u +v =− + + 2 , (8.2)
∂t ∂x ∂y ∂y Re ∂x2 ∂y
∂2p ∂2p
 
∂u ∂v ∂v ∂u
+ 2 =2 − . (8.3)
∂x2 ∂y ∂x ∂y ∂x ∂y
When using collocated grids, all three equations are approximated at the same
points in the grid for their respective dependent variables u(x, y), v(x, y), and
p(x, y).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 313 / 454

Grids and Grid Generation Staggered vs. Collocated Grids

Alternatively, using a staggered grid we approximate each primitive variable and


its associated equation as follows.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 314 / 454
Grids and Grid Generation Staggered vs. Collocated Grids

For example, let us consider how each of the terms in the x-momentum equation
(8.1) are approximated (using central differences) on a staggered grid

∂u ui+1,j − ui−1,j
u = ui,j + O(∆x2 ),
∂x 2∆x
∂u 1 ui,j+1 − ui,j−1
v = (vi,j + vi+1,j + vi,j−1 + vi+1,j−1 ) + O(∆y 2 ),
∂y 4 2∆y

∂p pi+1,j − pi,j
= + O(∆x2 ),
∂x ∆x

∂2u ui+1,j − 2ui,j + ui−1,j


= + O(∆x2 ),
∂x2 ∆x 2

∂2u ui,j+1 − 2ui,j + ui,j−1


= + O(∆y 2 ).
∂y 2 ∆y 2

The terms in equations (8.2) and (8.3) are treated in a similar manner at their
respective points while being approximated at their respective locations in the grid.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 315 / 454

Grids and Grid Generation Staggered vs. Collocated Grids

Note that special treatment is required at the boundaries.

Therefore, either grid points at which u or v are computed or specified do not


coincide with the boundary. Specifically, the points where u are taken do not
coincide with the y = 0 boundary, and v points do not coincide with the x = 0
boundary in the above example.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 316 / 454
Grids and Grid Generation Staggered vs. Collocated Grids

Advantages of staggered grid:


1 Helps avoid oscillations in pressure that can occur in some methods applied
to incompressible Navier-Stokes.
2 “Ensures” conservation of kinetic energy.
Advantages of collocated grid:
1 Easier to understand and program, i.e. ui,j , vi,j and pi,j are all approximated
at the same location.
2 Easier to implement boundary conditions, particularly for complex boundaries.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 317 / 454

Grids and Grid Generation Uniform vs. Non-Uniform Grids

Outline

8 Grids and Grid Generation


Introduction
Staggered vs. Collocated Grids
Uniform vs. Non-Uniform Grids
Grid Generation
Algebraic Grid Generation
Elliptic Grid Generation
Variational Grid Generation

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 318 / 454
Grids and Grid Generation Uniform vs. Non-Uniform Grids

In many flows, e.g. those involving boundary layers, the solution has local regions
of intense gradients.

Therefore, a fine grid is necessary to resolve the flow near boundaries, but the
same resolution is not necessary in the remainder of the domain. As a result, a
uniform grid would waste computational resources where they are not needed.
Alternatively, a non-uniform grid would allow us to refine the grid where it is
needed.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 319 / 454

Grids and Grid Generation Uniform vs. Non-Uniform Grids

Let us obtain the finite-difference approximations using Taylor series as before, but
without assuming all ∆x’s are equal. For example, consider the first-derivative
term ∂φ/∂x.

Applying Taylor series at xi−1 and xi+1 with ∆xi 6= ∆xi+1 and solving for ∂φ/∂x
leads to
∆x2i+1 − ∆x2i ∆x3i+1 + ∆x3i
 2   3 
∂φ φi+1 − φi−1 ∂ φ ∂ φ
= − − +· · · .
∂x ∆xi + ∆xi+1 2(∆xi + ∆xi+1 ) ∂x2 i 6(∆xi + ∆xi+1 ) ∂x3 i

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 320 / 454
Grids and Grid Generation Uniform vs. Non-Uniform Grids

If the grid is uniform, i.e. ∆xi = ∆xi+1 , then the second term vanishes, and the
approximation reduces to the usual O(∆x2 )-accurate central difference
approximation for the first derivative. However, for a non-uniform grid, the
truncation error is only O(∆x).
We could restore second-order accuracy by using an appropriate approximation to
∂ 2 φ/∂x2 i in the second term, which results in

∂φ φi+1 ∆x2i − φi−1 ∆x2i+1 + φi (∆x2i+1 − ∆x2i )


= + O(∆x2 ).
∂x ∆xi+1 ∆xi (∆xi + ∆xi+1 )

As one can imagine, this gets very complicated, and it is difficult to ensure
consistent accuracy for all approximations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 321 / 454

Grids and Grid Generation Grid Generation

Outline

8 Grids and Grid Generation


Introduction
Staggered vs. Collocated Grids
Uniform vs. Non-Uniform Grids
Grid Generation
Algebraic Grid Generation
Elliptic Grid Generation
Variational Grid Generation

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 322 / 454
Grids and Grid Generation Grid Generation

Rather than using non-uniform grids, it is often advantageous to transform the


problem to a computational domain on which a uniform mesh may be used. Such
a transformation generally has one or both of the following objectives:
1 Transform a complex physical domain into a simple, e.g. rectangular or
circular, computational domain.
2 Cluster grid points in regions of the physical domain where the solution varies
rapidly, i.e. large gradients occur.
We consider transformations of the form

(x, y) ⇐⇒ (ξ, η),

where (x, y) are the variables in the physical domain, and (ξ, η) are the variables
in the computational domain such that ξ = ξ(x, y) and η = η(x, y). To transform
back to the physical domain, we regard x = x(ξ, η) and y = y(ξ, η).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 323 / 454

Grids and Grid Generation Grid Generation

From the chain rule, the transformation laws are


∂ ∂ξ ∂ ∂η ∂
= + , (8.4)
∂x ∂x ∂ξ ∂x ∂η
∂ ∂ξ ∂ ∂η ∂
= + , (8.5)
∂y ∂y ∂ξ ∂y ∂η
where ξx = ∂ξ/∂x, ξy = ∂ξ/∂y, ηx = ∂η/∂x, and ηy = ∂η/∂y are called the
metrics of the transformation.
The Jacobian of the transformation is

ξx ηx ξx ξy
J = = = ξx ηy − ξy ηx . (8.6)
ξy η y η x ηy

Properties of the transformation:


1 Mapping must be one-to-one, which requires that J 6= 0.
2 We desire a smooth grid distribution; therefore, the metrics should be
smoothly varying throughout the domain.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 324 / 454
Grids and Grid Generation Grid Generation

3 We may want to enforce orthogonality of the grid (see Fletcher pp. 97-100).

Methods of obtaining the mapping:


1 Conformal mapping
See figure 1 for an example of a Schwarz-Christoffel transformation.
Difficult to extend to 3-D.
For details, recall MMAE 501.
2 Algebraic transformation.
3 Elliptic grid generation, i.e. solution of partial differential equation.
4 Variational methods.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 325 / 454

Grids and Grid Generation Grid Generation

Algebraic Grid Generation


Consider, for example, the flow in a diffuser:

The equation for the upper (sloped) boundary is

H2 − H1
yb = f (x) = H1 + x.
L

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 326 / 454
Grids and Grid Generation Grid Generation

We seek to transform this physical domain into a rectangular computational


domain through the algebraic tranformation
y
ξ = x, η= . (8.7)
f (x)

We then utilize a uniform grid in the computational domain (ξ, η).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 327 / 454

Grids and Grid Generation Grid Generation

The inverse transformation is then


 
H2 − H1
x = ξ, y = f (x)η = H1 + ξ η, (8.8)
L

which produces a stretched grid in the physical domain.


Let us check the metrics of this transformation
∂ξ
= 1,
∂x
∂ξ
= 0,
∂y
 H −H 
∂η f 0 (x) f 0 (ξ) 2
L
1
η
= −y 2 =− η=− ,
∂x f (x) f (ξ) H1 + H2 −H
L
1
ξ

∂η 1 1
= = .
∂y f (x) H1 + 2 −H
H
L
1
ξ

See figure 2 for plots of these metrics, which show that they are smooth.
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 328 / 454
Grids and Grid Generation Grid Generation

Let us transform the governing equations using the transformation laws (8.4) and
(8.5)
∂ ∂ξ ∂ ∂η ∂ ∂ f 0 (ξ) ∂
= + = −η ,
∂x ∂x ∂ξ ∂x ∂η ∂ξ f (ξ) ∂η

∂ ∂ξ ∂ ∂η ∂ 1 ∂
= + = .
∂y ∂y ∂ξ ∂y ∂η f (ξ) ∂η
For example, consider the convection term

f 0 (ξ) ∂u
 
∂u ∂u
u(x, y) = u(ξ, η) −η .
∂x ∂ξ f (ξ) ∂η

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 329 / 454

Grids and Grid Generation Grid Generation

Let us now consider another scenario in which we have a semi-infinite


boundary-layer flow above a surface.

We have introduced the boundary-layer variable Y such that

y = Re−1/2 Y, 0 ≤ Y ≤ ∞.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 330 / 454
Grids and Grid Generation Grid Generation

We want a transformation that maps the semi-infinite domain into a finite one
and clusters points near the surface. One possibility is
 
2 Y
ξ = x, η = tan−1 . (8.9)
π a

This maps Y = 0 to η = 0 and Y = ∞ to η = 1, and reducing a concentrates


more grid points near the surface at y = 0 in the physical domain when using a
uniform grid in the computational (ξ, η) domain.
For an illustration, see figure 3 in which transformation (8.9) is used in the
x-direction and a stretching transformation similar to that for the diffuser is used
in the y-direction.
The transformation laws for the transformation (8.9) are

∂ ∂
= ,
∂x ∂ξ

∂ ∂ξ ∂ ∂η ∂ ∂
= + = Γ(η) ,
∂Y ∂Y ∂ξ ∂Y ∂η ∂η

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 331 / 454

Grids and Grid Generation Grid Generation

where
1
Γ(η) = [1 + cos(πη)] .
πa
Then
∂2 ∂2
 
∂ ∂ ∂
2
= Γ(η) Γ(η) = Γ (η) 2 + Γ(η)Γ0 (η) .
2
∂Y ∂η ∂η ∂η ∂η
Notes:
1 Algebraic methods move the complexity, i.e. complex boundaries and/or
non-uniform grids, to the equations themselves.
Physical domain ⇒ “simple” equations; complex geometry and grid.
Computational domain ⇒ simple geometry and grid; “complex” equations.
2 Computational overhead is typically relatively small for algebraic methods, i.e.
there are no additional equations to solve.
3 It is easy to cluster grid points in the desired regions of the domain; however,
it is necessary to know where to cluster the grid points a priori.
4 Must choose the algebraic transformation ahead of time, i.e. before solving
the problem (cf. variational grid generation).
5 It is difficult to handle complex geometries.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 332 / 454
Grids and Grid Generation Grid Generation

Elliptic Grid Generation


Consider the potential flow past a cylinder:

The velocity potential and streamfunction both satisfy the Laplace equation

∂2φ ∂2φ ∂2ψ ∂2ψ


+ 2 = 0, + = 0. (8.10)
∂x2 ∂y ∂x2 ∂y 2
Observe that the streamlines and isopotential lines would make a good grid on
which to obtain a numerical solution, i.e. φ → ξ, ψ → η.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 333 / 454

Grids and Grid Generation Grid Generation

Thus, an appropriate solution to

∂2ξ ∂2ξ ∂2η ∂2η


+ = 0, + = 0, (8.11)
∂x2 ∂y 2 ∂x2 ∂y 2
would provide a good mapping on which to base the computational grid, i.e. a
uniform grid in (ξ, η) produces the above grid in physical space.
There are two approaches to obtaining the solution to (8.11):
1) Grid parameters are governed by the Laplace equation, i.e. they are harmonic;
therefore, we could utilize complex variables:

z = x + iy, ζ = ξ + iη.

This approach uses conformal mapping (see Fletcher, vol. II, pp. 89-96) and
is good for two-dimensional flows in certain types of geometries.
2) Solve a boundary-value problem to generate the grid, i.e. elliptic grid
generation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 334 / 454
Grids and Grid Generation Grid Generation

In order to control grid clustering, known functions involving sources and sinks
may be added to the right-hand-side of equations (8.11)

∂2ξ ∂2ξ ∂2η ∂2η


+ = P (x, y), + = Q(x, y), (8.12)
∂x2 ∂y 2 ∂x2 ∂y 2

where P and Q contain exponential functions. Note that equations (8.12) are for
ξ = ξ(x, y) and η = η(x, y).
We want to solve (8.12) in the computational domain (ξ, η) to obtain the grid
transformations x = x(ξ, η) and y = y(ξ, η). In addition, we must transform the
governing equation(s), e.g. Navier-Stokes, to the computational domain.
Therefore, we seek the transformation laws for (x, y) → (ξ, η).
For
ξ = ξ(x, y), η = η(x, y),
the total differentials are
∂ξ ∂ξ ∂η ∂η
dξ = dx + dy, dη = dx + dy,
∂x ∂y ∂x ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 335 / 454

Grids and Grid Generation Grid Generation

or in matrix form     
dξ ξ ξy dx
= x . (8.13)
dη ηx ηy dy
Similarly, for the inverse transformation from (ξ, η) to (x, y), we have that
    
dx xξ xη dξ
= .
dy yξ yη dη

We can solve the latter expression for [dξ dη]T by multiplying by the inverse
   −1  
dξ xξ xη dx
=
dη yξ yη dy
 
yη −xη
−yξ xξ
 
dx
=
xξ xη dy

yξ yη
    
dξ 1 yη −xη dx
=
dη J −yξ xξ dy

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 336 / 454
Grids and Grid Generation Grid Generation

Comparing with equation (8.13), we see that


   
ξx ξy 1 yη −xη
= .
ηx ηy J −yξ xξ

Thus, the transformation metrics are


∂ξ 1 ∂y ∂ξ 1 ∂x
= , =− ,
∂x J ∂η ∂y J ∂η
∂η 1 ∂y ∂η 1 ∂x
=− , = .
∂x J ∂ξ ∂y J ∂ξ

Substituting into the transformation laws (8.4) and (8.5) give


 
∂ ∂ξ ∂ ∂η ∂ 1 ∂y ∂ ∂y ∂
= + = − ,
∂x ∂x ∂ξ ∂x ∂η J ∂η ∂ξ ∂ξ ∂η
 
∂ ∂η ∂ ∂ξ ∂ 1 ∂x ∂ ∂x ∂
= + = − .
∂y ∂y ∂η ∂y ∂ξ J ∂ξ ∂η ∂η ∂ξ

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 337 / 454

Grids and Grid Generation Grid Generation

From these we obtain the second derivative transformation laws as follows


 2 2  2 2
∂2 ∂2ξ ∂ ∂2η ∂ ∂ξ ∂ ∂η ∂ ∂η ∂ξ ∂ 2
= + + + + 2 , (8.14)
∂x2 ∂x2 ∂ξ ∂x2 ∂η ∂x ∂ξ 2 ∂x ∂η 2 ∂x ∂x ∂η∂ξ

and
2 2
∂2 ∂2ξ ∂ ∂2η ∂ ∂2 ∂2 ∂η ∂ξ ∂ 2
 
∂ξ ∂η
= + 2 + + +2 , (8.15)
∂y 2 ∂y 2 ∂ξ ∂y ∂η ∂y ∂ξ 2 ∂y ∂η 2 ∂y ∂y ∂η∂ξ

Application of equations (8.14) and (8.15) to equations (8.12) gives

∂2x ∂2x ∂2x


 
1 ∂x ∂x
g22 2 − 2g12 + g11 2 = − 2 P +Q ,
∂ξ ∂ξ∂η ∂η J ∂ξ ∂η
(8.16)
∂2y ∂2y ∂2y
 
1 ∂y ∂y
g22 2 − 2g12 + g11 2 = − 2 P +Q ,
∂ξ ∂ξ∂η ∂η J ∂ξ ∂η

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 338 / 454
Grids and Grid Generation Grid Generation

where  2  2
∂x ∂y
g11 = + ,
∂ξ ∂ξ
∂x ∂x ∂y ∂y
g12 = + ,
∂ξ ∂η ∂ξ ∂η
 2  2
∂x ∂y
g22 = + .
∂η ∂η
Note that the coefficients in the equations are defined such that

g11 g12
g = = J 2 , (g12 = g21 ).
g21 g22

For boundary conditions to apply to equations (8.16), specify either:


1) Grid point locations (x, y) on boundaries.
2) Grid line slopes at the boundaries.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 339 / 454

Grids and Grid Generation Grid Generation

Therefore, we have a coupled set of nonlinear elliptic equations with Dirichlet or


Neumann boundary conditions. We solve (GS, SOR, ADI, multigrid, etc...)
equations (8.16) on a uniform grid in (ξ, η) to obtain the inverse transformation
x = x(ξ, η), y = y(ξ, η).
Notes:
1) Provides smoother grids than algebraic transformations due to use of
diffusion-type equations to generate grid.
→ Effects of discontinuities at boundaries are smoothed out in interior.
2) Can enforce grid orthogonality.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 340 / 454
Grids and Grid Generation Grid Generation

3) Can treat multiply-connected regions (see figure 5).

4) Can extend to three-dimensional geometries (cf. conformal mapping).


5) Difficult to choose P (ξ, η) and Q(ξ, η) for desired clustering. Generally, one
must use trial-and-error until something that “looks good” is obtained.
6) Large computational times are required to produce the grid, i.e. on the same
order as the time required to solve the governing equations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 341 / 454

Grids and Grid Generation Grid Generation

7) References:
For more elliptic grid generation options, see Knupp, P. and Steinberg, S.,
“Fundamentals of Grid Generation,” CRC Press (1994), who consider
smoothness, Winslow and TTM methods.
Thompson, J. F., Warsi, Z. U. A. and Mastin, C. W., “Numerical Grid
Generation - Foundation and Applications,” North Holland (1985).
8) The ultimate in grid generation is adaptive grid methods, in which the grid
“adapts” to local features of the solution as it is computed; see, for example,
variational grid generation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 342 / 454
Grids and Grid Generation Grid Generation

Variational Grid Generation


Advantages:
Provides a more intuitive and mathematically formal basis for grid generation.
Generally formulated to produce the optimal, i.e. ”best,” grid in a least
squares sense; therefore, it eliminates the trial-and-error necessary in elliptic
grid generation.
One-Dimensional:
We determine functionals for which the stationary function(s) gives a
transformation in which the grid spacing ∆xi is proportional to a weight function
φ(ξi+1/2 ).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 343 / 454

Grids and Grid Generation Grid Generation

Physical domain: a ≤ x ≤ b

Computational domain: 0 ≤ ξ ≤ 1

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 344 / 454
Grids and Grid Generation Grid Generation

Consider the following discrete functional


I I
X (xi+1 − xi )2 X ∆x2i
S= = ,
i=1
2φi+1/2 i=1
2φ i+1/2

where xi = x(ξi ), xi+1 = x(ξi+1 ) and φi+1/2 = (φ(ξi ) + φ(ξi+1 ))/2. For a given
weight function φ(ξ), we want to minimize S subject to the end conditions
x1 = a, xI+1 = b.
Dividing the above expression by ∆ξ (= constant) gives
I  2
S X ∆xi ∆ξ
= .
∆ξ i=1
∆ξ 2φi+1/2

Taking ∆ξ → 0 gives the continuous functional


1
(∂x/∂ξ)2 1 x2ξ
Z Z
1 1
I[x(ξ)] = dξ = dξ, (8.17)
2 0 φ(ξ) 2 0 φ

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 345 / 454

Grids and Grid Generation Grid Generation

with the end conditions


x(0) = a, x(1) = b.
This is known as the weighted length functional. We seek the grid distribution
x(ξ) that minimizes the functional I[x(ξ)].
With the integrand of the functional being

1 x2ξ
F = F (ξ, x, xξ ) = ,
2 φ
Euler’s equation is given by
 
∂F d ∂F
− =0
∂x dξ ∂xξ
 
d xξ
− =0 (8.18)
dξ φ
φξ
∴ xξξ − xξ = 0, (φ > 0), (8.19)
φ

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 346 / 454
Grids and Grid Generation Grid Generation

with boundary conditions


x(0) = a, x(1) = b.
In order to see what this means for the grid spacing, let us consider equation
(8.18), which is  

= 0.
φ ξ
Integrating (with integration constant C) and multiplying by φ(ξ) gives

dx
= Cφ(ξ).

Writing this expression in discrete form yields
 
∆xi ξi+1 + ξi
=Cφ .
∆ξ 2

Therefore, we see that this requires that ∆xi be proportional to φi+1/2 , where the
proportionality constant is C∆ξ.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 347 / 454

Grids and Grid Generation Grid Generation

Note that the weight function φ(ξ) has been expressed in the computational
domain, and it is not necessarily clear how it should be chosen. Conceptually, we
prefer to think in terms of physical weight functions, say w(x). In that case, the
grid spacing is determined by a physical variable or one of its derivatives giving
rise to feature-adaptive grid generation.

Let us set the grid spacing to be proportional to [w(x)]2 , i.e.


  2
xi+1 + xi
∆xi = K w .
2

Therefore, we take φ = w2 > 0 in equation (8.17) giving the physical weight


length functional
1 1 x2ξ
Z
I[x(ξ)] = dξ, (8.20)
2 0 w2 (x)
with x(0) = a, x(1) = b.
In this case, the integrand of the functional is

1 x2ξ
F = F (ξ, x, xξ ) = .
2 w2 (x)
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 348 / 454
Grids and Grid Generation Grid Generation

Therefore,
∂F wx ∂F xξ
= − 3 x2ξ , = 2,
∂x w ∂xξ w
and Euler’s equation gives
wx 2 d  xξ 
− 3 xξ − =0
w dξ w2
wx 2 xξξ wx xξ
x + − 2 xξ = 0
w3 ξ w2 w3
wx 2
xξξ − x = 0.
w ξ
By the chain rule
d dξ d 1 d 1
= = ⇒ wx = wξ .
dx dx dξ xξ dξ xξ

Thus, the Euler equation is



xξξ − xξ = 0, (w > 0), (8.21)
w

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 349 / 454

Grids and Grid Generation Grid Generation

which is of the same form as (8.19) (this is why we set φ = w2 ), except that now
w(x) is a physical weight function. The boundary conditions are

x(0) = a, x(1) = b.

One-Dimensional Illustration:
Consider the one-dimensional, steady convection-diffusion equation

uxx − cux = 0, a ≤ x ≤ b,
(8.22)
u(a) = 0, u(b) = 1,

where c is the convection speed. Increasing c produces an increasingly thin


boundary layer near x = b. The exact solution to (8.22) is

ec(x−a) − 1
u(x) = c(b−a) .
e −1

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 350 / 454
Grids and Grid Generation Grid Generation

The governing equation (8.22) and the grid equation must be transformed into
the computational domain. In one dimension, the transformation laws are
d dξ d 1 d
= = ,
dx dx dξ xξ dξ
d2 1 d2
 
1 d 1 d xξξ d
= = 2 2− 3 .
dx2 xξ dξ xξ dξ xξ dξ xξ dξ

Transforming the governing equation (8.22) to computational coordinates gives

1 d2 u xξξ du 1 du
− − c = 0,
x2ξ dξ 2 x3ξ dξ xξ dξ
or  
xξξ
uξξ − + cxξ uξ = 0, (8.23)

with boundary conditions
u(0) = 0, u(1) = 1.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 351 / 454

Grids and Grid Generation Grid Generation

To solve (8.23) for the velocity u(ξ) requires the grid x(ξ). Here we will use the
grid equation (8.21) (corresponding to the physical weighted-length functional
(8.20))

xξξ − xξ = 0, (8.24)
w
with the boundary conditions x(0) = a, x(1) = b.
→ How to choose w(x)?
A common choice for a feature-adaptive weight function is based on the gradient
of the dependent variable of the form
1
w(x) = p , 0 < w ≤ 1, (8.25)
1 + 2 u2x

where  is a parameter to be chosen. Observe that

ux small ⇒ w ≈ 1,
1
ux large ⇒ w≈ , (|ux | ↑ ⇒ w ↓ ⇒ ∆x ↓).
|ux |

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 352 / 454
Grids and Grid Generation Grid Generation

For use in the grid equation (8.24), we need the weight function (8.25) in terms of
the computational coordinate ξ:
1
w (x(ξ)) = q . (8.26)
2 2
1+ u
x2ξ ξ

Taking the derivative


 
2 x 2 xξξ
2 x2ξ uξ uξξ − 2 xξξ
3 uξ 2 uξ xξ uξ − uξξ
ξ
wξ = − = 3/2 .
2 
2 2
3/2 x2ξ  2 2
1 + x2 uξ 1 + x2 uξ
ξ ξ

Therefore, the coefficient in equation (8.24) is


 
xξξ
wξ uξ xξ uξ − uξξ
= 2 . (8.27)
w x2ξ + 2 u2ξ

Note: A uniform grid in the physical domain results if  = 0 (w = 1, wξ = 0), i.e.


x(ξ) = (b − a)ξ + a. Thus, increasing  increases the influence of the
transformation.
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 353 / 454

Grids and Grid Generation Grid Generation

Numerical Procedure:
The governing equation (8.23) and the grid equation (8.24) (with (8.27)
expressed in the computational ξ-plane are

uξξ − X(ξ)uξ = 0, u(0) = 0, u(1) = 1, (8.28)

xξξ − W (ξ)xξ = 0, x(0) = a, x(1) = b, (8.29)


respectively, where
xξξ
X(ξ)
+ cxξ , =


W (ξ) = .
w
Recall that the last expression is given by equation (8.27). Thus, we have
two coupled second-order linear ordinary differential equations. Note that all
of the grid transformation information is encapsulated in the X(ξ) coefficient
in the governing equation, and all the weight function information is
contained in the W (ξ) coefficient in the grid equation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 354 / 454
Grids and Grid Generation Grid Generation

Consider equation (8.28) approximated on the computational grid using


central differences
ui+1 − 2ui + ui−1 ui+1 − ui−1
2
− Xi = 0.
∆ξ 2∆ξ
Thus, we have the tridiagonal problem

Ai ui−1 + Bi ui + Ci ui+1 = Di , i = 2, . . . , I,

where
∆ξ
Ai = 1+ Xi ,
2
Bi = −2,
∆ξ
Ci = 1− Xi ,
2
Di = 0.
We obtain a similar tridiagional problem for the grid equation (8.29).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 355 / 454

Grids and Grid Generation Grid Generation

See the Mathematica notebook ’GridGen.nb’


Two-Dimensional:
The two-dimensional analog to the one-dimensional weighted-length functional
(8.17) is
Z Z " #
1 1 1 x2ξ + yξ2 x2η + yη2
I[x, y] = + dξdη,
2 0 0 φ(ξ, η) ψ(ξ, η)

where we now have two weight functions φ(ξ, η) > 0 and ψ(ξ, η) > 0. This
produces a grid for which the lengths of the coordinate lines are proportional to
the weight functions, i.e.
√ q
g11 = x2ξ + yξ2 = K1 φ(ξ, η),
√ q
g22 = x2η + yη2 = K2 ψ(ξ, η).

Recall that g11 and g22 are defined in section 8.4.2.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 356 / 454
Grids and Grid Generation Grid Generation

The Euler equations are then


   
xξ xη
+ = 0,
φ ξ ψ η
   
yξ yη
+ = 0.
φ ξ ψ η

Notes:
1) If φ = ψ = c, then the Euler equations are Laplace equations

xξξ + xηη = 0, yξξ + yηη = 0;

cf. elliptic grid generation.


2) The weight functions control the interior grid locations in a much more
formal manner than the P and Q forcing functions in elliptic grid generation.
That is, the local grid spacing is directly proportional to the specified weight
functions, which may be related directly to the physical solution.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 357 / 454

Grids and Grid Generation Grid Generation

Other functionals are also possible:


Area Functional:
The area of each cell, which is related to the Jacobian J, is proportional to a
weight function φ(ξ, η) > 0, i.e.
1 1
J2
Z Z
1
∴ IA [x, y] = dξdη.
2 0 0 φ

Euler equations:    
Jxη Jxξ
− = 0,
φ ξ φ η
   
Jyη Jyξ
− = 0.
φ ξ φ η

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 358 / 454
Grids and Grid Generation Grid Generation

Orthogonality Functional:
The grid is orthogonal if g12 = xξ xη + yξ yη = 0; therefore, the orthogonality
functional is
1 1 1 2
Z Z
∴ IO [x, y] = g dξdη,
2 0 0 12
such that g12 is minimized in a least squares sense (without a weight
function).
Euler equations:
(g12 xη )ξ + (g12 xξ )η = 0,
(g12 yη )ξ + (g12 yξ )η = 0.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 359 / 454

Grids and Grid Generation Grid Generation

Combination Functionals:
We can form combinations of the above functionals. For example, consider
the area-orthogonality functional
1 1 1 1
J 2 + g12
2
Z Z Z Z
1 1 g11 g22
IAO [x, y] = dξdη, = dξdη,
2 0 0 φ 2 0 0 φ

where again the fact that J 2 = g = g11 g22 − g12


2
has been used.
This results in a grid such that the average of the area and orthogonality
functionals are minimized. One criteria could be emphasized over the other
by including weight coefficients in the respective terms in the functional.
Euler equations:    
g22 xξ g11 xη
+ = 0,
φ ξ φ η
   
g22 yξ g11 yη
+ = 0.
φ ξ φ η

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 360 / 454
Grids and Grid Generation Grid Generation

Notes:
1) Just as in the one-dimensional case, it is generally preferable to define weight
functions in the physical domain, i.e. w(x, y), rather than in the
computational domain, i.e. φ(ξ, η).
2) The grid x(ξ, η), y(ξ, η) can be obtained by solving the Euler equations (most
common) or the variational form directly.
3) All of the two-dimensional functionals above have been written in the form
Z 1 Z 1
1
I[x, y] = F (ξ, η, x, y, xξ , yξ , xη , yη )dξdη,
2 0 0

in order to determine the mappings x(ξ, η) and y(ξ, η). Alternatively,


functionals may be written to determine ξ(x, y) and η(x, y), i.e.
Z 1 Z 1
1
I[ξ, η] = F (x, y, ξ, η, ξx , ηx , ξy , ηy )dxdy.
2 0 0

These are called contravariant functionals (see Knupps and Steinberg section
8.5 and chapter 11).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 361 / 454

Grids and Grid Generation Grid Generation

4) For unsteady problems additional terms are required in the governing


equations due to movement of the grid in the physical domain (the
computational grid (ξ, η) remains fixed and uniform), i.e.

ξ = ξ(x, y, t), η = η(x, y, t).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 362 / 454
Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

Outline

9 Hydrodynamic Stability and the Eigenproblem


1-D, Unsteady Diffusion Illustration
Exact Solution
Numerical Solution
Numerical Solution of the Eigenproblem
Similarity Transformation
QR Method to Obtain Eigenvalues and Eigenvectors
Plane Rotations
Arnoldi Method
Hydrodynamic Stability
Linearized Navier-Stokes Equations
Local Normal Mode Analysis
Numerical Solution of the Orr-Sommerfeld Equation
Example: Plane-Poiseuille Flow

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 363 / 454

Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

Consider the one-dimensional, unsteady diffusion equation

∂u ∂2u
= α 2, 0 ≤ x ≤ `, (9.1)
∂t ∂x
with the boundary conditions

u(0, t) = 0, u(`, t) = 0, (9.2)

and the initial condition


u(x, 0) = f (x). (9.3)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 364 / 454
Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

Exact Solution
Let us begin by using the method of separation of variables to obtain an exact
solution for equation (9.1). We separate the variables according to

u(x, t) = φ(x)ψ(t). (9.4)

Substituting into equation (9.1) gives

dψ d2 φ
φ = αψ 2 .
dt dx
Moving everything depending upon t to the left-hand-side and everything
depending upon x to the right-hand-side, this becomes

1 dψ 1 d2 φ
= = λ = −µ2 .
αψ dt φ dx2
Because the x and t dependence can be separated in this way, both sides of the
equation must be equal to a constant, say λ. Positive and zero λ produce only the

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 365 / 454

Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

trivial solution for the boundary conditions given, so we consider the case where
λ = −µ2 < 0.
Therefore, the partial differential equation (9.1) is converted into two ordinary
differential equations
d2 φ
2
+ µ2 φ = 0, (9.5)
dx

+ αµ2 ψ = 0, (9.6)
dt
each of which are differential eigenproblems. The solution to equation (9.5) is

φ(x) = c1 cos(µx) + c2 sin(µx). (9.7)

The boundary condition u(0, t) = 0 requires that φ(0) = 0, which requires that
c1 = 0. From the boundary condition u(`, t) = 0, we must have φ(`) = 0, which
requires that sin(µn `) = 0. Therefore,

µn = , n = 1, 2, 3, . . . ,
`

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 366 / 454
Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

and the eigenvalues are


n2 π 2
λn = =− 2 . −µ2n
`
Letting c2 = 1, we have the spatial eigenfunction
 nπ 
φn (x) = sin x , n = 1, 2, 3, . . . (9.8)
`
The solution to (9.6) is
2
ψn (t) = cn e−αµn t , n = 1, 2, 3, . . . (9.9)

Then the eigenfunctions are


∞ ∞ ∞
αn2 π 2
X X    nπ  X
u(x, t) = un (x, t) = φn (x)ψn (t) = cn exp − 2 t sin x .
n=1 n=1 n=1
` `
(9.10)
The constants cn are determined by application of the initial condition (9.3), for
which
X∞  nπ 
u(x, 0) = cn sin x = f (x),
n=1
`
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 367 / 454

Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

or

X
cn φn (x) = f (x).
n=1

Taking the inner product of φm (x) with both sides, the only non-vanishing term
(due to orthogonality of the eigenfunctions φn (x)) occurs when m = n, giving

cn ||φn (x)||2 = hf (x), φn (x)i ;

Therefore,
`
hf (x), φn (x)i
Z
2  nπ 
cn = 2
= f (x) sin x dx, n = 1, 2, 3, . . . , (9.11)
||φn (x)|| ` 0 `

which are the Fourier sine coefficients of f (x). Thus, the exact solution to
(9.1)–(9.3) is given by equation (9.10) with (9.11).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 368 / 454
Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

Numerical Solution
Now let us consider solving the differential eigenproblem (9.5), with boundary
conditions (9.2) numerically. Using central differences, the differential equation
becomes
φi+1 − 2φi + φi−1
= λφi .
(∆x)2
Thus, for i = 2, . . . , I (φ1 = 0, φI+1 = 0), in matrix form we have
    
−2 1 0 ··· 0 0 φ2 φ2
 1 −2 1 ··· 0 0   φ3 
   φ3 
  
0 1 −2 · · · 0 0  φ4   φ4 
2 
..   ..  = (∆x) λ  ..  ,
   
 .. .. .. .. ..
 .
 . . . . . 
 . 
   . 
 
0 0 0 · · · −2 1  φI−1  φI−1 
0 0 0 ··· 1 −2 φI φI
or
Aφ = λ̄φ. (9.12)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 369 / 454

Hydrodynamic Stability and the Eigenproblem 1-D, Unsteady Diffusion Illustration

This is an algebraic eigenproblem for the eigenvalues λ̄ = (∆x)2 λ and the


eigenvectors φ, which are discrete approximations of the continuous eigenfunctions
(9.8).
Because A is tridiagonal with constants along each diagonal, we have a closed
form expression for the eigenvalues (see section 3.5). In general, of course, this is
not the case, and the eigenvalues (and eigenvectors) must be determined
numerically for a large matrix A.
See Mathematica notebook “1Ddiff.nb” for a comparison of the eigenvalues
obtained numerically as described here versus the exact eigenvalues from the
previous section.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 370 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Outline

9 Hydrodynamic Stability and the Eigenproblem


1-D, Unsteady Diffusion Illustration
Exact Solution
Numerical Solution
Numerical Solution of the Eigenproblem
Similarity Transformation
QR Method to Obtain Eigenvalues and Eigenvectors
Plane Rotations
Arnoldi Method
Hydrodynamic Stability
Linearized Navier-Stokes Equations
Local Normal Mode Analysis
Numerical Solution of the Orr-Sommerfeld Equation
Example: Plane-Poiseuille Flow

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 371 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

The standard method for numerically determining the eigenvalues and eigenvectors
of a matrix is based on QR decomposition, which entails performing a series of
similarity transformations. This is the approach used by the built-in Mathematica
and Matlab functions Eigenvalues[]/ Eigenvectors[] and eig(), respectively.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 372 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Similarity Transformation
Consider the eigenproblem
Ax = λx, (9.13)
where A is a real, square matrix. Suppose that Q is an orthogonal matrix such
that Q−1 = QT . Let us consider the transformation

B = QT AQ. (9.14)

Postmultiplying both sides by QT x leads to

BQT x = QT AQQT x,
= QT Ax,
= QT λx,
BQT x = λQT x.

Defining y = QT x, this can be written as

By = λy, (9.15)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 373 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

which is an eigenproblem for the matrix B defined by the transformation (9.14).


Note that the eigenproblems (9.13) and (9.15) have the same eigenvalues λ;
therefore, we call equation (9.14) a similarity transformation because A and
B = QT AQ have the same eigenvalues. This is the case because Q is orthogonal.
The eigenvectors of A, i.e. x, and B, i.e. y, are related by

y = QT x (x = Qy) . (9.16)

If in addition to being real and square, A is symmetric such that A = AT ,


observe that

BT = [QT AQ]T = [Q]T [A]T [QT ]T = QT AQ = B.

Therefore, if A is symmetric, B is symmetric as well when Q is orthogonal.


In summary, for A real and symmetric, the similarity transformation (9.14)
preserves the eigenvalues and symmetry of A, and the eigenvectors are related by
(9.16).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 374 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

QR Method to Obtain Eigenvalues and Eigenvectors


Having determined the properties of similarity transformations, we now turn our
attention to the iterative QR method for finding the eigenvalues and eigenvectors
of a matrix A, which requires such transformations.
Consider A real and symmetric, i.e. A = AT . A QR decomposition exists such
that
A = QR,
where Q is an orthogonal matrix, and R is an upper (“right”) triangular matrix.
Letting
A0 = A, Q0 = Q, R0 = R,
the QR decomposition of the given matrix is

A 0 = Q0 R0 . (9.17)

Let us form the product (note order)

A 1 = R0 Q0 . (9.18)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 375 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Because Q0 is orthogonal, premultiplying equation (9.17) by Q−1 T


0 = Q0 gives

QT0 A0 = QT0 Q0 R0 = R0 .

Therefore, substituting for R0 in (9.18) we may determine A1 from

A1 = QT0 A0 Q0 , (9.19)

which is a similarity transformation. That is, taking R0 Q0 is equivalent to the


similarity transformation (9.19), and A1 has the same eigenvalues as A0 = A.
Thus, generalizing equation (9.18) we have

Ak+1 = Rk Qk , k = 0, 1, 2, . . . (9.20)

where all A1 , A2 , . . . , Ak , . . . are similar to A0 = A. Not only do they all have


the same eigenvalues, the similarity transformations maintain the same structure
as A, e.g. tridiagonal.
It can be shown (not easily) that the sequence of similar matrices

A0 , A1 , A2 , . . .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 376 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

gets progressively closer, i.e. converges, to a diagonal or upper triangular matrix if


A is symmetric or non-symmetric, respectively. In either case, the eigenvalues of
A (and A1 , A2 , . . .) are on the main diagonal in increasing order by absolute
magnitude.
But how do we determine the Q and R matrices for each iteration?
⇒ Plane rotations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 377 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Plane Rotations
Consider the n × n transformation matrix P comprised of the identity matrix with
only four elements changed in the pth and q th rows and columns according to

Ppp = Pqq = c, Ppq = s, Pqp = −s,

where c = cos φ and s = sin φ. That is,


 
1
 .. 

 . 


 1 


 c 0 ··· 0 s 


 0 1 0 

P= .. .. ..
.
 
 . . . 

 0 1 0 


 −s 0 ··· 0 c 


 1 

 .. 
 . 
1
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 378 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Observe the effect of transforming a n-D vector x according to the tranformation

y = Px. (9.21)

where xT = [x1 x2 · · · xp · · · xq · · · xn ]. Then


 
x1
 x2 
 .. 
 
 . 
 
 yp 
y = Px =  .  ,
 
 .. 
 
 yq 
 
 . 
 .. 
xn

where the only two elements that are altered are

yp = cxp + sxq , (9.22)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 379 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

yq = −sxp + cxq . (9.23)


For example, consider the case for n = 2, i.e. p = 1, q = 2

y1 = cx1 + sx2 ,

y2 = −sx1 + cx2 ,
or     
y1 cos φ sin φ x1
= .
y2 − sin φ cos φ x2
This transformation rotates the vector x through an angle φ to obtain y. Note
that y = PT x rotates the vector x through an angle −φ.
Thus, in the general n-D case (9.21), P rotates the vector x through an angle φ
in the xp xq -plane.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 380 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Notes:
1 The transformation matrix P is orthogonal, i.e. PT = P−1 .
2 We can generalize to rotate a set of vectors, i.e. a matrix, by taking

Y = PX.

3 The angle φ may be chosen with one of several objectives in mind. For
example,
i) To zero all elements below (or to the right of) a specified element, e.g.
yT = [y1 y2 · · · yj 0 · · · 0].
Householder transformation (reflection) – efficient for dense matrices.
ii) To zero a single element, e.g. yp or yq (see equations (9.22) and (9.23)).
Givens transformation (rotation) – efficient for sparse, structured (e.g. banded)
matrices.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 381 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

We can imagine a series of Givens or Householder transformations that reduce the


matrix A to a matrix that is upper triangular, which is the R matrix in a QR
decomposition.
Thus, if m projections are required to produce an upper triangular matrix, R is
given by
R = Pm · · · P2 P1 A. (9.24)
Because A = QR, i.e. R = QT A, the orthogonal matrix Q is then obtained from

QT = Pm · · · P2 P1 .

Taking the transpose leads to

Q = PT1 PT2 · · · PTm . (9.25)

See Mathematica notebook “QRmethod.nb” for an illustration of how QR


decomposition is used in an iterative algorithm to obtain the eigenvalues of a
matrix.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 382 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Notes:
1 The QR decomposition (9.24) and (9.25) is obtained from a series of plane
(Givens or Householder) rotations.
2 Givens transformations are most efficient for large, sparse, structured
matrices.
→ Configure to only zero elements that are not already zero.
3 There is a “fast Givens transformation” for which the P matrices are not
orthogonal, but the QR decompositions can be obtained two times faster
than in the standard Givens transformation illustrated in “QRmethod.nb.”
4 Convergence of the iterative QR method may be accelerated using shifting
(see, for example, Numerical Recipes, section 11.3).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 383 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

5 The order of operations for the QR method per iteration are as follows:
Dense matrix → O(n3 ) ⇒ Very expensive.
Hessenberg matrix → O(n2 )
Tridiagonal matrix → O(n).
Thus, the most efficient procedure is as follows:
i) Transform A to a similar tridiagonal or Hessenberg form if A is symmetric or
non-symmetric, respectively.
→ This is done using a series of similarity transformations based on Householder
rotations for dense matrices or Givens rotations for sparse matrices.
ii) Use iterative QR method to obtain eigenvalues of tridiagonal or Hessenberg
matrix.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 384 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Arnoldi Method
The Arnoldi method has been developed to treat situations in which we only need
a small number of eigenvalues of a large sparse matrix:
1 The iterative QR method described in the previous section is the general
approach used to obtain the full spectrum of eigenvalues of a dense matrix.
2 As we saw in the 1-D unsteady diffusion example, and as we will see when we
evaluate hydrodynamic stability, we often seek the eigenvalues of large sparse
matrices.
3 In addition, we often do not require the full spectrum of eigenvalues in
stability problems as we only seek the “least stable mode.”
⇒ We would like an efficient algorithm that determines a subset of the full
sprectrum of eigenvalues (and possibly eigenvectors) of a sparse matrix.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 385 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

Suppose we seek the largest k eigenvalues (by magnitude) of the large sparse
n × n matrix A, where k  n. Given an arbitrary n-D vector q0 , we define the
Krylov subspace by

Kk (A, q0 ) = span q0 , Aq0 , A2 q0 , . . . , Ak−1 q0 ,




which has dimension k and is a subspace of Rn .


The Arnoldi method is based on constructing an orthonormal basis, e.g. using
Gram-Schmidt, of the Krylov subspace Kk that can be used to project a general
n × n matrix A onto the k-D Krylov subspace Kk (A, q0 ).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 386 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

We form the orthonormal projection matrix Q using the following step-by-step


(non-iterative) method that produces a Hessenberg matrix H whose eigenvalues
approximate the largest k eigenvalues of A:
1 Specify starting Arnoldi vector q0 .
2 Normalize: q1 = q0 /||q0 ||.
3 Set Q = q1 .
4 Do i = 2, k
i) Multiply qi = Aqi−1 .
ii) Orthogonalize qi against q1 , q2 , . . . , qi−1 .
iii) Append qi to Q.
iv) Form the Hessenberg matrix H = QT AQ.
v) Determine the eigenvalues of H.
5 End Do

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 387 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

At each step i = 2, . . . , k:
→ An n × i orthonormal matrix Q is produced that forms an orthonormal basis
for the Krylov subspace Ki (A, q0 ).
→ Using the projection matrix Q, we transform A to produce an i × i
Hessenberg matrix H (or tridiagonal for symmetric A), which is an
orthogonal projection of A onto the Krylov subspace Ki .
→ The eigenvalues of H, sometimes called the Ritz eigenvalues, approximate
the largest i eigenvalues of A.
The approximations of the eigenvalues improve as each step is incorporated, and
we obtain the approximation of one additional eigenvalue.
Notes:
1 Because k  n, we only require the determination of eigenvalues of
Hessenberg matrices that or no larger than k × k as opposed to the original
n × n matrix A.
2 Although the outcome of each step depends upon the starting Arnoldi vector
q0 used, the procedure converges to the correct eigenvalues of matrix A.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 388 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

3 The more sparse the matrix A is, the smaller k can be to obtain a good
approximation of the largest k eigenvalues of A.
4 When applied to symmetric matrices, the Arnoldi method reduces to the
Lanczos method.
5 A shift and invert approach can be incorporated to determine the k
eigenvalues close to a specified part of the spectrum rather than that with
the largest magnitude.
For example, it can be designed to determine the k eigenvalues with the
largest real or imaginary part.
6 When seeking a set of eigenvalues in a particular portion of the full spectrum,
it is desirable that the starting Arnoldi vector q0 be in (or ‘nearly’ in) the
subspace spanned by the eigenvectors corresponding to the sought after
eigenvalues.
As the Arnoldi method progresses, we get better approximations of the desired
eigenvectors that can then be used to form a more desirable starting vector.
This is known as the implicitly restarted Arnoldi method and is based on the
implicitly-shifted QR decomposition method.
Restarting also reduces storage requirements by keeping k small.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 389 / 454

Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

7 The Arnoldi method may also be adapted to solve linear systems of


equations; this is called the generalized minimal residual (GMRES) method.
8 The Arnoldi method can be designed to apply to the generalized eigenproblem

Ax = λBx.

The generalized eigenproblem will be encountered in hydrodynamic stability.


It also arises in structural design problems in which A is called the stiffness
matrix, and B is called the mass matrix.
Application of Arnoldi’s method to the generalized eigenproblem requires that
B be positive definite, i.e. have all positive eigenvalues.
9 Many have standardized on the Arnoldi method as implemented in ARPACK
(http://www.caam.rice.edu/software/ARPACK/).
ARPACK was developed at Rice University in the mid 1990’s, first as a Fortran
77 library of subroutines, and subsequently it has been implemented as
ARPACK++ for C++ .
ARPACK has been implemented in Matlab via the eigs() function, where the
‘s’ denotes ‘sparse.’
In addition, it has been implemented in Mathematica, where one includes the
option ‘Method → Arnoldi’ in the Eigenvalues[] function.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 390 / 454
Hydrodynamic Stability and the Eigenproblem Numerical Solution of the Eigenproblem

The Arnoldi method is illustrated in more detail in the Mathematica notebook


“Arnoldi.nb.”
References:
Arnoldi, W. (1951) Q. Appl. Math. 9, 17.
Nayar, N. & Ortega, J. M. (1993) “Computation of Selected Eigenvalues of
Generalized Eigenvalue Problems.” J. Comput. Phys. 108, pp. 8–14.
Saad, Y. Iterative Methods for Sparse Linear Systems SIAM, Philadelphia
(2003).
Radke, R. A Matlab Implementation of the Implicitly Restarted Arnoldi
Method for Solving Large-Scale Eigenvalue Problems, MS Thesis, Rice
University (1996).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 391 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Outline

9 Hydrodynamic Stability and the Eigenproblem


1-D, Unsteady Diffusion Illustration
Exact Solution
Numerical Solution
Numerical Solution of the Eigenproblem
Similarity Transformation
QR Method to Obtain Eigenvalues and Eigenvectors
Plane Rotations
Arnoldi Method
Hydrodynamic Stability
Linearized Navier-Stokes Equations
Local Normal Mode Analysis
Numerical Solution of the Orr-Sommerfeld Equation
Example: Plane-Poiseuille Flow

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 392 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Linearized Navier-Stokes Equations


Consider the nondimensional Navier-Stokes equations for 2-D, incompressible flow
∂u ∂v
+ = 0, (9.26)
∂x ∂y

1 ∂2u ∂2u
 
∂u ∂u ∂u ∂p
+u +v =− + + 2 , (9.27)
∂t ∂x ∂y ∂x Re ∂x2 ∂y
 2
∂2v

∂v ∂v ∂v ∂p 1 ∂ v
+u +v =− + + 2 . (9.28)
∂t ∂x ∂y ∂y Re ∂x2 ∂y
We denote the solution to (9.26)–(9.28), i.e. the base flow, by u0 (x, y, t),
v0 (x, y, t) and p0 (x, y, t), and seek the behavior of small perturbations to this
base flow.
⇒ If the amplitude of the small perturbations grow, the flow is
hydrodynamically unstable.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 393 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Two possibilities may be considered:


1 Temporal analysis – amplitude of spatial perturbation (e.g. wavy wall)
grows/decays with time.
⇒ Absolutely unstable/stable.
2 Spatial analysis – amplitude of temporal perturbation (e.g. vibrating ribbon)
grows/decays in space.
⇒ Convectively unstable/stable.
For infinitesimally small perturbations (  1), the flow may be decomposed as
follows
u(x, y, t) = u0 (x, y, t) + û(x, y, t),
v(x, y, t) = v0 (x, y, t) + v̂(x, y, t), (9.29)
p(x, y, t) = p0 (x, y, t) + p̂(x, y, t).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 394 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Substituting into (9.26)–(9.28) gives

∂u0 ∂ û ∂v0 ∂v̂


+ + + = 0,
∂x ∂x ∂y ∂y
   
∂u0 ∂ û ∂u0 ∂ û ∂u0 ∂ û
+ + (u0 + û) + + (v0 + v̂) +
∂t ∂t ∂x ∂x ∂y ∂y
 2 2 2 2

∂p0 ∂ p̂ 1 ∂ u0 ∂ û ∂ u0 ∂ û
=− − + +  + +  ,
∂x ∂x Re ∂x2 ∂x2 ∂y 2 ∂y 2
   
∂v0 ∂v̂ ∂v0 ∂v̂ ∂v0 ∂v̂
+ + (u0 + û) + + (v0 + v̂) +
∂t ∂t ∂x ∂x ∂y ∂y
 2
∂ 2 v̂ ∂ 2 v0 ∂ 2 v̂

∂p0 ∂ p̂ 1 ∂ v0
=− − + +  + +  .
∂y ∂y Re ∂x2 ∂x2 ∂y 2 ∂y 2
As expected, the O(1) terms are simply the Navier-Stokes equations (9.26)–(9.28)
for the base flow u0 , v0 and p0 .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 395 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

The O() terms for the disturbance flow are

∂ û ∂v̂
+ = 0, (9.30)
∂x ∂y

1 ∂ 2 û ∂ 2 û
 
∂ û ∂ û ∂ û ∂u0 ∂u0 ∂ p̂
+ u0 + v0 + û + v̂ = − + + 2 , (9.31)
∂t ∂x ∂y ∂x ∂y ∂x Re ∂x2 ∂y
 2
∂ 2 v̂

∂v̂ ∂v̂ ∂v̂ ∂v0 ∂v0 ∂ p̂ 1 ∂ v̂
+ u0 + v0 + û + v̂ = − + + 2 . (9.32)
∂t ∂x ∂y ∂x ∂y ∂y Re ∂x2 ∂y
Because  is small, we neglect O(2 ) terms. Thus, the evolution of the
disturbances are governed by the linearized Navier-Stokes (LNS) equations
(9.30)–(9.32), where the base flow is known.
⇒ Linear Stability Theory

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 396 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

In principle, we could impose a perturbation û, v̂, p̂ at any time ti and track its
evolution in time and space to determine if the flow is stable to the imposed
perturbation. To fully characterize the stability of the base flow, however, would
require many calculations of the LNS equations with different perturbation
“shapes” imposed at different times.
We can formulate a more manageable stability problem by doing one or both of
the following:
1 Consider simplified base flows.
2 Impose “well-behaved” perturbations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 397 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Local Normal Mode Analysis


Classical linear stability analysis takes advantage of both simplifications mentioned
above:
1 Base flow:
i) Steady
∂u0 ∂v0
⇒ = = 0. (9.33)
∂t ∂t
ii) Parallel (e.g. Poiseuille flow)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 398 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

⇒ u0 = u0 (y), v0 = 0, p0 = p0 (x). (9.34)


2 Perturbations – we impose normal modes for û, v̂, p̂ of the form (temporal
analysis):
û(x, y, t) = u1 (y)ei(αx−αct) ,
v̂(x, y, t) = v1 (y)ei(αx−αct) , (9.35)
p̂(x, y, t) = p1 (y)ei(αx−αct) ,
where α is the wavenumber, which is real, and c = cr + ici is the complex
wavespeed. Note that the wavelength of the disturbance is proportional to
1/α.
Notes:
1 It is understood that we take the real parts of (9.35) (or add complex
conjugate).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 399 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

2 Consider the disturbances, e.g.


 
û(x, y, t) = Re u1 (y)ei(αx−αct) , c = cr + ici
 
= Re u1 (y)eαci t eαi(x−cr t) ,
= Re [u1 (y)eαci t {cos[α(x − cr t)] + i sin[α(x − cr t)]}] ,
û(x, y, t) = u1 (y)eαci t cos[α(x − cr t)]

⇒ Sine wave with wavenumber α and phase velocity cr , i.e. normal mode.
If ci > 0, the amplitude of the perturbation grows unbounded as t → ∞ with
growth rate αci .
3 Because equations (9.30)–(9.32) are linear, each normal mode with
wavenumber α may be considered independently of one another (cf. von
Neumann numerical stability analysis).
⇒ For a given mode α, we are looking for the eigenvalue (wavespeed) with the
fastest growth rate.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 400 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

4 This is regarded as a “local” analysis because the stability of only one


velocity profile (9.34), i.e. at a single streamwise location, is considered at a
time due to the parallel-flow assumption.
5 In some cases, the parallel flow assumption can be justified on formal grounds
if the wavelength is such that 1/α  L, where L is a typical streamwise
length scale in the flow.
6 For the temporal analysis considered here, α is real and c is complex. For a
spatial analysis, α is complex and c is real.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 401 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

For steady, parallel base flow (9.33) and (9.34), the Navier-Stokes equations
(9.26)–(9.28) reduces to (from equation (9.27))

d2 u0 ∂p0
= Re , (9.36)
dy 2 ∂x

where Rep00 (x) is a constant for Poiseuille flow. The disturbance equations
(9.30)–(9.32) become
∂ û ∂v̂
+ = 0, (9.37)
∂x ∂y
1 ∂ 2 û ∂ 2 û
 
∂ û ∂ û ∂u0 ∂ p̂
+ u0 + v̂ = − + + 2 , (9.38)
∂t ∂x ∂y ∂x Re ∂x2 ∂y
1 ∂ 2 v̂ ∂ 2 v̂
 
∂v̂ ∂v̂ ∂ p̂
+ u0 =− + + 2 . (9.39)
∂t ∂x ∂y Re ∂x2 ∂y

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 402 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Substitution of the normal modes (9.35) into the disturbance equations


(9.37)–(9.39) leads to
αiu1 + v10 = 0, (9.40)
1 00
αi(u0 − c)u1 + u00 v1 = −αip1 + (u1 − α2 u1 ), (9.41)
Re
1 00
αi(u0 − c)v1 = −p01 +
(v1 − α2 v1 ), (9.42)
Re
where primes denote differentiation with respect to y.
Solving equation (9.40) for u1 and substituting into equation (9.41) results in
1 1 000
−(u0 − c)v10 + u00 v1 = −αip1 − (v1 − α2 v10 ), (9.43)
αi Re
leaving equations (9.42) and (9.43) for v1 (y) and p1 (y). Solving equation (9.43)
for p1 (y), differentiating and substituting into equation (9.42) leads to
1 1 0000
(u0 − c)(v100 − α2 v1 ) − u000 v1 = (v − 2α2 v100 + α4 v1 ), (9.44)
αi Re 1
which is the Orr-Sommerfeld equation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 403 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Notes:
1 For a given base flow u0 (y), Reynolds number Re, and wavenumber α, the
Orr-Sommerfeld equation is a differential eigenproblem of the form

L1 v1 = cL2 v1 ,

where the wavespeeds c are the (complex) eigenvalues, the disturbance


velocities v1 (y) are the eigenfunctions, and L1 and L2 are differential
operators.
2 The Orr-Sommerfeld equation applies for steady, parallel, viscous flow
perturbed by infinitesimally small normal modes, which is a local analysis.
3 For inviscid flow, i.e. as Re → ∞, the Orr-Sommerfeld equation reduces to
the Rayleigh equation

(u0 − c)(v100 − α2 v1 ) − u000 v1 = 0.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 404 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

4 For non-parallel flows, a significantly more involved global stability analysis is


required in which the base flow is 2-D, i.e.

u0 = u0 (x, y), v0 = v0 (x, y), p0 = p0 (x, y),

and the perturbations are of the form

û(x, y, t) = u1 (x, y)e−ict ,


v̂(x, y, t) = v1 (x, y)e−ict ,
p̂(x, y, t) = p1 (x, y)e−ict ,

in place of equation (9.35).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 405 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Numerical Solution of the Orr-Sommerfeld Equation


Recall that the continuous differential eigenproblem (9.44), i.e. the
Orr-Sommerfeld equation, is of the form

L1 v1 = cL2 v1 .

We seek a corresponding discrete, i.e. algebraic, generalized eigenproblem of the


form
M(α, Re)v = cN(α)v, (9.45)
where we have dropped the subscript on v. Let us rewrite the Orr-Sommerfeld
equation in the form
 
i 0000
v + Pj v 00 + Qj v = c v 00 − α2 v ,
 
(9.46)
αRe

where
2αi
P (yj ) = u0 (yj ) − ,
Re
α3 i
Q(yj ) = − α2 u0 (yj ) − u000 (yj ).
Re
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 406 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

To discretize equation (9.46) using central differences, we take


d2 v vj+1 − 2vj + vj−1
2
= 2
+ O(∆y 2 ),
dy (∆y)
d4 v vj+2 − 4vj+1 + 6vj − 4vj−1 + vj−2
4
= 4
+ O(∆y 2 ).
dy (∆y)
Substituting these approximations into equation (9.46) and collecting terms leads
to the difference equation
 
[Cvj−2 + Bj vj−1 + Aj vj + Bj vj+1 + Cvj+2 ] = c B̄vj−1 + Āvj + B̄vj+1 ,
(9.47)
where
6i
+ (∆y)2 (∆y)2 Qj − 2Pj ,
 
Aj =
αRe
4i
Bj = − + (∆y)2 Pj ,
αRe
i
C = ,
αRe
 
Ā = −(∆y)2 2 + α2 (∆y)2 ,
B̄ = (∆y)2 .
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 407 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Because the Orr-Sommerfeld equation is 4th -order, we need two boundary


conditions at each boundary. For solid surfaces at y = a, b, we set

v = v 0 = 0, at y = a, b. (9.48)

Because v is known at y = a, b, i.e. v1 = vJ+1 = 0, the unknowns are


vj , j = 2, . . . , J.
Applying equation (9.47) at j = 2, we have
   
0 0
Cv0 + B2v
> v
1 + A2 v2 + B2 v3 + Cv4 = c B̄1 + Āv2 + B̄v3 ,
>

and from v 0 = 0 at y = a (j = 1)
v2 − v0
=0 ⇒ v 0 = v2 .
2∆y
Substituting into the difference equation for j = 2 results in
 
[(C + A2 )v2 + B2 v3 + Cv4 ] = c Āv2 + B̄v3 . (9.49)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 408 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Similarly, for j = J

[CvJ−2 + BJ vJ−1 + (C + AJ )vJ ] = c [BvJ−1 + AvJ ] . (9.50)

Also for j = 3, J − 1, we have v1 = vJ+1 = 0. Therefore, the matrices in the


algebraic form of the eigenproblem (9.45) are
 
C + A2 B 2 C 0 0 ··· 0 0 0 0
 B3
 A3 B 3 C 0 · · · 0 0 0 0  
 C B 4 A 4 B 4 C · · · 0 0 0 0 
M(α, Re) =  . ,
 
. . . . . . . . .
 .. .. .. .. .. . . .. .. .. .. 
 
 0 0 0 0 0 · · · C BJ−1 AJ−1 BJ−1 
0 0 0 0 0 ··· 0 C BJ C + AJ
 
Ā B̄ 0 0 ··· 0 0 0
B̄
 Ā B̄ 0 ··· 0 0 0 
0 B̄ Ā B̄ ··· 0 0 0
N(α) =  . ..  ,
 
.. .. .. .. .. ..
 .. . . . . . . .
 
0 0 0 0 ··· B̄ Ā B̄ 
0 0 0 0 ··· 0 B̄ Ā
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 409 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

and the eigenvector is vT = [v2 v3 · · · vJ ]. Thus, M is pentadiagonal, and N is


tridiagonal.
The (large) generalized eigenproblem (9.45) must be solved to obtain the complex
wavespeeds c, i.e. the eigenvalues, and the discretized eigenfunctions v(y) for a
given α and Re.
Notes:
1 The J − 1 eigenvalues obtained approximate the infinity of eigenvalues of the
continuous differential eigenproblem (9.44), i.e. the Orr-Sommerfeld
equation.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 410 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

2 The least stable mode, i.e. that with the fastest growth rate αci , is given by
αmax(ci ).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 411 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

3 A marginal stability curve may be obtained by determining max(ci ) for a


range of Re and α and plotting the max(ci ) = 0 contour:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 412 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Methods of Solution:
1 Convert the generalized eigenproblem (9.45) to a regular eigenproblem by
multiplying both sides by the inverse of N. That is, find the eigenvalues from
−1
N M − cI = 0.

This requires inverting a large matrix, and although M and N are typically
sparse and banded, N−1 M is a full, dense matrix.
⇒ This would require use of a general approach, such as the iterative QR
method, to determine the eigenvalues of a large, dense matrix.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 413 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

2 In order to avoid solving the large matrix problem that results from the BVP,
traditionally the shooting method for IVPs has been used.

→ This approach avoided the need to find the eigenvalues of large matrices in the
days when computers were not capable of such large calculations.
→ In addition, it allowed for use of well-developed algorithms for IVPs.
→ However, this is like “using a hammer to drive a screw.”

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 414 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

3 Solve the generalized eigenproblem

Mv = cNv,

where M and N are large, sparse matrices, for the least stable mode. In
addition to the fact that the matrices are sparse, in stability contexts such as
this, we only need the least stable mode, not the entire spectrum of
eigenvalues. Recall that the least stable mode is that with the largest
imaginary part.
→ Currently, the state-of-the-art in such situations is the Arnoldi method
discussed in the last section.
Note that N must be positive definite for use in the Arnoldi method. That is,
it must have all positive eigenvalues. In our case, this requires us to take the
negatives of the matrices M and N as defined above.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 415 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

Example: Plane-Poiseuille Flow


As an illustration, let us consider stability of Plane-Poiseuille flow, i.e.
pressure-driven flow in a channel.

Such a flow is parallel; therefore, the base flow is a solution of equation (9.36).
The solution is a parabolic velocity profile given by

u0 (y) = y(2 − y), 0 ≤ y ≤ 2. (9.51)

Note that the base flow is independent of the Reynolds number.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 416 / 454
Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

See the Mathematica notebook “OS Psvll.nb” for a solution of the


Orr-Sommerfeld equation using the approach outlined in the previous section.
→ This notebook calculates the complex wavespeeds, i.e. the eigenvalues, for a
given wavenumber α and Reynolds number Re in order to evaluate stability.
→ Recall that a flow is unstable if the imaginary part of one of the discrete
eigenvalues is positive, i.e. max(ci ) > 0. The growth rate of the instability is
then α max(ci ).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 417 / 454

Hydrodynamic Stability and the Eigenproblem Hydrodynamic Stability

By performing a large number of such calculations for a range of wavenumbers


and Reynolds numbers, we can plot the marginal stability curve for
plane-Poiseuille flow. This shows the curve in α-Re parameter space for which
max(ci ) = 0 delineating the regions of parameter space in which the flow is stable
and unstable to normal-mode perturbations.
1.10

1.05

1.00

0.95

0.90

0.85

0.80
5000 6000 7000 8000 9000 10 000

Thus, the critical Reynolds number is approximately Rec = 5, 800.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 418 / 454
Numerical Modeling of Turbulent Flows Introduction

Outline

10 Numerical Modeling of Turbulent Flows


Introduction
Direct Numerical Simulation
Large-Eddy Simulation
Reynolds Averaged Navier-Stokes

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 419 / 454

Numerical Modeling of Turbulent Flows Introduction

Turbulent flows are characterized by the following:


They are highly unsteady ⇒ Seemingly random fluctuations of velocity about
a mean.
They are inherently three-dimensional due to velocity fluctuations (even if the
mean flow is two-dimensional).
They involve a wide range of temporal and spatial scales.
Thus, turbulent flows are very difficult to model numerically.
There are three basic approaches (given in increasing degree of approximation):
1) Direct Numerical Simulation (DNS):
Navier-Stokes equations are solved for all scales. ⇒ Very fine grids are
required.
2) Large Eddy Simulation (LES):
Recognize that only the large-scale motions depend on the macroscopic
features of the particular flow, and that the small-scale features are similar for
certain classes of turbulent flows. ⇒ Solve numerically for large scales and
model small scales.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 420 / 454
Numerical Modeling of Turbulent Flows Introduction

3) Reynolds Averaged Navier-Stokes (RANS):


Solve numerically the time averaged Navier-Stokes equations for the mean
flow.
Requires turbulence models for closure of the governing equations.
Summarizing:

DNS ⇒ no turbulence modeled.


LES ⇒ some, i.e. small-scale, turbulence modeled.
RANS ⇒ all turbulence modeled.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 421 / 454

Numerical Modeling of Turbulent Flows Direct Numerical Simulation

Outline

10 Numerical Modeling of Turbulent Flows


Introduction
Direct Numerical Simulation
Large-Eddy Simulation
Reynolds Averaged Navier-Stokes

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 422 / 454
Numerical Modeling of Turbulent Flows Direct Numerical Simulation

In DNS all scales must be resolved.


Consider the following definitions:

 = rate of dissipation of turbulent kinetic energy


 3 1/4
ν
lK = = Kolmogorov scale (size of the smallest turbulent eddies)

L = length scale of large eddies
ReL = Reynolds number based on magnitude of
velocity fluctuations and L, e.g. ReL ≈ 0.01Re.

Here, Re is the macroscopic Reynolds number.


The grid size must be smaller than lK in order to resolve all of the scales;
therefore,
L 3/4
N∼ ∼ ReL ,
lK
where N is the number of grid points in each direction.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 423 / 454

Numerical Modeling of Turbulent Flows Direct Numerical Simulation

Note that as ReL ↑ ⇒ lK ↓ ⇒ N ↑; therefore, DNS is only practical for small to


moderate Re.
Notes:
1) DNS is only possible for moderate Re flows in simple geometries (with
current computational resources). This gives rise to the early, i.e. 1990s,
”turbulence-in-a-box” simulations.
2) DNS produces very detailed information about the flow (see DNS figures):
⇒ Can be used to evaluate turbulence statistics, e.g. to compare with
experiments.
⇒ Gives qualitative knowledge of flow (coherent structures, etc...).
⇒ Used to develop turbulence models for similar flows.
3) Spectral methods are popular for DNS because highly accurate solutions can
be obtained.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 424 / 454
Numerical Modeling of Turbulent Flows Large-Eddy Simulation

Outline

10 Numerical Modeling of Turbulent Flows


Introduction
Direct Numerical Simulation
Large-Eddy Simulation
Reynolds Averaged Navier-Stokes

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 425 / 454

Numerical Modeling of Turbulent Flows Large-Eddy Simulation

In LES we filter the velocity and pressure fields so they only contain large-scale
components, i.e. local average of complete fields:
Z
ūi (xi ) = G(x, x0 )ui (x0 )dx0 , (10.1)

where G(x, x0 ) is the filter kernel having an associated length scale ∆:


Eddies > ∆ ⇒ large eddies (computed).
Eddies < ∆ ⇒ small turbulent eddies (modeled).
That is,
lK ≤ modeled < ∆ < computed ≤ L
Numerical grid: lK < ∆xi < ∆, i.e. smaller than ∆, but not as small as lK (cf.
DNS).
Apply filter (10.1) to Navier-Stokes equations with

ui (xi , t) = ūi (xi , t) + u0i (xi , t), p(xi , t) = p̄(xi , t) + p0 (xi , t),

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 426 / 454
Numerical Modeling of Turbulent Flows Large-Eddy Simulation

where ūi (xi , t) is the resolvable scale velocity (computed), and u0i (xi , t) is the
subgrid scale (SGS) velocity.
The filtered Navier-Stokes equations are (in tensor notation)

∂ ūi
= 0,
∂xi
(10.2)
∂ ūi ∂ 1 ∂ p̄
+ ui uj = − + ν∇2 ūi .
∂t ∂xj ρ ∂xi

Thus, we have a set of equations for large-scale quantities, but

ui uj = ūi ūj + ūi u0j + u0i ūj + u0i u0j . (10.3)

The first term on the right-hand-side is computed from the resolved scales, and
the remaining terms are τij = SGS Reynolds stress and must be modeled, i.e. they
contain u0i and u0j .
Therefore, the subgrid scale (SGS) model must specify τij as a function of the
resolvable variables (ūi , ūj ), and it provides for energy transfer between the
resolvable scales and the SGS.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 427 / 454

Numerical Modeling of Turbulent Flows Large-Eddy Simulation

The earliest and most common SGS model is the Smagorinsky model, which is an
eddy viscosity model (effective viscosity due to small-scale turbulent motion).
Thus, the SGS Reynolds stress τij increases transport and dissipation.

1
τij = τkk δij + 2µt S̄ij , (10.4)
3
where  
1 ∂ ūi ∂ ūj
S̄ij = + = strain-rate for resolved field
2 ∂xj ∂xi
µt = CS2 ρ∆2 |S̄|.
Here, CS is the model parameter, which must be specified, and |S̄| = (S̄ij S̄ij )1/2 .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 428 / 454
Numerical Modeling of Turbulent Flows Large-Eddy Simulation

Notes:
1) The Smagorinsky model only accounts for energy transfer from large to small
scales.

2) The Smagorinsky model does not work well near boundaries (eddy viscosity is
much smaller and flow is more anisotropic).
3) Points (1) and (2) can be improved upon with dynamic SGS models:
Allow model parameter CS to vary with space and time, i.e. it is computed
from the resolvable flow field.
Automatically adjusts SGS parameter for anisotropic flow and flow near walls.
Allows for backscatter (µt < 0), which accounts for energy transferred from
small scales to large scales.
Active area of ongoing research.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 429 / 454

Numerical Modeling of Turbulent Flows Large-Eddy Simulation

4) LES is more economical computationally than DNS; therefore, it can be


applied to more complex flows. Consider, for example, the flow over a wall
mounted block (Re = 3, 200 and grid: 240 × 128 × 128) and combustor flows
(see figures) and note the level of detail obtained for the flow.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 430 / 454
Numerical Modeling of Turbulent Flows Reynolds Averaged Navier-Stokes

Outline

10 Numerical Modeling of Turbulent Flows


Introduction
Direct Numerical Simulation
Large-Eddy Simulation
Reynolds Averaged Navier-Stokes

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 431 / 454

Numerical Modeling of Turbulent Flows Reynolds Averaged Navier-Stokes

In RANS we compute only the mean flow and model all of the turbulence:

u(xi , t) = ū(xi ) + u0 (xi , t), (10.5)

where ū(xi ) is the mean flow, and u0 (xi , t) are the turbulent fluctuations (not the
same as u0 in LES).
To obtain the mean flow, we use Reynolds averaging:
Steady mean flow ⇒ time average Navier-Stokes equations
Z T
1
ū(xi ) = lim u(xi , t)dt, (10.6)
T →∞ T 0

where T = averaging interval.


Unsteady mean flow ⇒ ensemble average.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 432 / 454
Numerical Modeling of Turbulent Flows Reynolds Averaged Navier-Stokes

For incompressible flow, the Reynolds averaged Navier-Stokes equations are

∂ ū ∂v̄
+ = 0,
∂x ∂y
     
∂ ū ∂ ū ∂ ū ∂ p̄ ∂ ∂ ū ∂ ∂ ū
ρ + ū + v̄ =− + µ − ρu0 u0 + µ − ρu0 v 0 ,
∂t ∂x ∂y ∂x ∂x ∂x ∂y ∂y
     
∂v̄ ∂v̄ ∂v̄ ∂ p̄ ∂ ∂v̄ 0 0
∂ ∂v̄ 0 0
ρ + ū + v̄ =− + µ − ρu v + µ − ρv v ,
∂t ∂x ∂y ∂y ∂x ∂x ∂y ∂y
(10.7)
0 0 0 0 0 0
where ρu u , ρu v , ρv v are the Reynolds stresses. Equations (10.7) are three
equations for five unkowns (ū, v̄, p̄, u0 , v 0 ); therefore, we have a closure problem.
Closure is achieved by relating the Reynolds stresses to the mean flow quantities
through a turbulence model.
k −  Turbulence Model:
The k −  model is the most common turbulence model used in applications, but
there are many others.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 433 / 454

Numerical Modeling of Turbulent Flows Reynolds Averaged Navier-Stokes

Define:
1 0 0
k = u u = turbulent kinetic energy,
2 i i
 0  0 
∂ui ∂ui
 = νT = rate of dissipation of turbulent energy.
∂xj ∂xj

Here, k and  are determined from a solution of the following coupled equations,
which are derived from Navier-Stokes,
   
Dk ∂ µT ∂k ∂ui ∂uj ∂ui
ρ = + µT + − ρ, (10.8)
Dt ∂xj σk ∂xj ∂xj ∂xi ∂xj

ρC2 2
   
D ∂ µT ∂ C1 µT  ∂ui ∂uj ∂ui
ρ = + + − . (10.9)
Dt ∂xj σ ∂xj k ∂xj ∂xi ∂xj k
The terms on the left-hand-side represent the transport of k and , and the first,
second and third terms on the right-hand-side represent diffusion, production and
dissipation, respectively, of k and .

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 434 / 454
Numerical Modeling of Turbulent Flows Reynolds Averaged Navier-Stokes

µT (xi ) is the local eddy viscosity and is related to k and  by

Cµ ρk 2
µT = . (10.10)

The eddy viscosity µT is used to relate the Reynolds stresses to the mean flow
quantities through
 
∂ ūi ∂ ūj 2
−ρu0i u0j = µT + − ρkδij . (10.11)
∂xj ∂xi 3

Notes:
1) Equations (10.8)–(10.10) involve five constants, Cµ , C1 , C2 , σk , σ , which
must be determined empirically.
2) Special treatment is necessary at solid boundaries; therefore, wall functions
that are usually based on the log-law are used for turbulent boundary layers.
3) Produces less detailed information about the flow than DNS and LES, but
RANS requires significantly more modest computational resources as
compared to DNS and LES. Thus, it is good for engineering applications, and
is used by most commercial CFD codes.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 435 / 454

Numerical Modeling of Turbulent Flows Reynolds Averaged Navier-Stokes

4) Most turbulence models (including k − ) have difficulties with accurately


predicting separation.
Consider the example of an impinging jet (see figures). Note the significant
differences between the results obtained with the two models used, both in terms
of the locations and magnitudes of turbulent kinetic energy.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 436 / 454
Parallel Computing Introduction

Outline

11 Parallel Computing
Introduction

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 437 / 454

Parallel Computing Introduction

Glossary:

Supercomputer – the fastest computers available at a particular time.

MIPS – Millions of Instructions Per Second; outdated measure of computer
performance (architecture dependent).

Floating point operation – an operation, e.g. addition, subtraction,
multiplication or division, on one or more floating point numbers (non
integers).

FLOPS – Floating Point Operations Per Second; current measure of
computer performance for numerical applications.
MFLOPS MegaFLOPS million (106 ) FLOPS
GFLOPS GigaFLOPS billion (109 ) FLOPS
TFLOPS TeraFLOPS trillion (1012 ) FLOPS
PFLOPS PetaFLOPS 1015 FLOPS

Serial processing – a computer code is executed line-by-line in sequence on
one processor.

Scalar processing – a processor performs calculations on a single data
element.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 438 / 454
Parallel Computing Introduction


Vector processing – a processor performs calculations on multiple data
elements, i.e. a vector, simultaneously.

Parallel processing – multiple processors (or cores) perform operations
simultaneously on different data elements.

Massively parallel – parallel computers involving thousands of processors.

MPP – Massively Parallel Processing.

SMP – Symmetric Multi-Processing; shared memory parallelism.

Shared memory parallelism – all of the processors share the same memory.

Distributed memory parallelism – all of the processors access their own
memory.

MPI – Message Passing Interface; most common library used for
inter-processor communication on distributed memory computers.

Cluster – a parallel computer comprised of commodity hardware, e.g. Beowulf
cluster (commodity hardware, Linux OS, and open source software).

SIMD – Single Instruction, Multiple Data; all processors perform the same
instruction on different data elements.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 439 / 454

Parallel Computing Introduction


MIMD – Multiple Instruction, Multiple Data; all processors perform different
instructions on different data elements.

Embarrassingly parallel – processors work in parallel with very little or no
communication between them, e.g. image processing and Seti@Home.

Coupled parallel – processors work in parallel but require significant
communication between them, e.g. CFD.

HPC – High Performance Computing.

Grid computing – parallel computing across a geographically distributed
network (e.g. the internet); analogous to the electric grid.

Multi-core CPUs – a single chip with multiple processors (cores).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 440 / 454
Parallel Computing Introduction

A Brief History of Supercomputers:


Generations of Supercomputer Architectures:

Note that each architecture requires its own approach to programming, and not
all algorithms are amenable to each.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 441 / 454

Parallel Computing Introduction

Milestones:
Year Supercomputer Peak Speed
1906 Babbage Analytical Engine 0.3 OPS
1946 ENIAC 50 kOPS
1964 CDC 6600 (Seymour Cray) 3 MFLOPS
1969 CDC 7600 36 MFLOPS
1976 Cray-1 (Seymour Cray) 250 MFLOPS
1981 CDC Cyber 205 400 MFLOPS
1983 Cray X-MP 941 MFLOPS
1985 Cray-2 3.9 GFLOPS
1985 Thinking Machines CM-2 (64k)
1989 Cray Y-MP
1993 Thinking Machines CM-5 65.5 GFLOPS
1993 Intel Paragon 143.4 GFLOPS
1996 Hitachi/Tsukuba CP-PACS (2k) 368.2 GFLOPS
1999 Intel ASCI Red (10k) 2.8 TFLOPS
2002 NEC Earth Simulator (5k) 35.9 TFLOPS
2005 IBM Blue Gene/L (131k) 280.6 TFLOPS
2008 IBM Roadrunner (130k) 1.105 PFLOPS
MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics
c 2010 K. W. Cassel 442 / 454
Parallel Computing Introduction

Milestones (cont’d):
1954 – Fortran (Formula Translation) developed
1970 – Unix developed at AT&T Bell Labs
Components of Computing:
1 Hardware – the computer (CPUs, memory, storage, network, etc.)
→ Assuming adequate memory, the computational time is primarily determined
by:
Sequential → CPU speed
Parallel → number of CPUs, CPU speed, inter-processor communication speed
and bandwidth, percent of code that is run in parallel (see Amdahl’s law).
2 Software – OS, compiler, libraries, program, etc.
3 Algorithms – the numerical method
→ Note that the best algorithms for serial computers often are not the best for
parallel computers (e.g. BLKTRI).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 443 / 454

Parallel Computing Introduction

Issues in Parallel Computing:


1 Data dependency
→ Can calculations be carried out independently of one another on separate
processors?
→ Examples:
Thomas algorithm ⇒ no.
Red-black Gauss-Siedel ⇒ yes.
2 Load balancing
→ Desire all processors to be utilized 100% of the time.
3 Scalability
→ Ability of a parallel code to maintain its speedup with increasing numbers of
processors and increasing problem size ⇒ see Amdahl’s Law.
4 Coarse vs. fine-grain parallelism
Fine-grain parallelism ⇒ parallelize the lower level structures of a program,
e.g. loops.
Coarse-grain parallelism ⇒ parallelize the higher level structures of a program,
e.g. equations.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 444 / 454
Parallel Computing Introduction

5 Architectures:
SIMD (e.g. CM-2) vs. MIMD
Shared memory (e.g. Cray, SGI, multicore CPUs, etc.):

Distributed memory (e.g. CM, clusters, etc.):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 445 / 454

Parallel Computing Introduction

Key Ingredients That Make Clusters Possible:


Hardware:
1) Fast, cheap PCs – inexpensive and widely available CPUs, memory, storage.
2) High-performance (high-bandwidth, low latency) LANs, e.g. Gigabit Ethernet,
Myrinet, Infiniband.
Software:
3) Open source OS, e.g. Linux – common platform on which to build tools
(previously, most supercomputing vendors had their own proprietary tools, e.g.
CM Fortran).
4) MPI – standardize message passing on various platforms (developed at ANL).

→ Ingredients (3) and (4) were not available during MPP (e.g. CM) heyday
in mid ’80s - mid ’90s. This led to the demise of the CM.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 446 / 454
Parallel Computing Introduction

Amdahl’s Law:
Measure of speedup S(N ) on N processors versus single processor, i.e. serial code.
Amdahl’s Law:
T (1) N
S(N ) = = ,
T (N ) N − (N − 1)F
where T (N ) is the time required using N processors, and F is the fraction of the
serial run time that the code spends in the parallel portion.
See Mathematica notebook ‘’AmdahlsLaw.nb.”
Notes:
1 If F = 1 (100% parallel) ⇒ S(N ) = N (linear speedup)
If F < 1 ⇒ S(N ) < N (sub-linear speedup)

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 447 / 454

Parallel Computing Introduction

2 Maximum speedup for given F with N → ∞:


N 1 1
lim S(N ) = lim = = ,
N →∞ N →∞ N − (N − 1)F 1−F Fs

where Fs = 1 − F is the fraction of time in the serial code spent doing the
serial portion.
e.g. F = 0.95 ⇒ Fs = 0.05 ⇒ S(∞) = 20
3 In practice, Fs (and F ) depends on the number of processors N and the size
of the problem n, i.e. the number of grid points.
Thus, for ideal scalability a parallel algorithm should be such that

Fs (N, n) → 0 (F → 1) as n → ∞

∴ S(N ) → N as n → ∞ ⇒ linear speedup


4 Parallel efficiency – fraction of total potential (linear) speedup achieved:

S(N )
E(N ) =
N

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 448 / 454
Parallel Computing Introduction

5 Illustrates importance of profiling – determining where the code spends most


of its time in order to know where to focus parallelization efforts.
6 Ignores overhead due to parallelization:

Inter-processor communication

Synchronization of data between processors

Load imbalances

Modified Amdahl’s Law:


Modified to account for parallel overhead.
Modified Amdahl’s Law:
N
S(N ) = ,
N − (N − 1)F + N 2 Fo

where Fo is the fraction of time spent due to parallel overhead.


See Mathematica notebook “AmdahlsLaw.nb.”

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 449 / 454

Parallel Computing Introduction

Top 500 Supercomputers (www.top500.org)


Compiled every six months based on LINPACK benchmark tests.

November 2005:

26th List: The TOP10


Rmax
Manufacturer Computer Installation Site Country Year #Proc
[TF/s]
BlueGene/L
1 IBM 280.6 DOE/NNSA/LLNL USA 2005 131072
eServer Blue Gene
BGW
2 IBM 91.29 IBM Thomas Watson USA 2005 40960
eServer Blue Gene
ASC Purple
3 IBM 63.39 DOE/NNSA/LLNL USA 2005 10240
eServer pSeries p575
4 Columbia
SGI 51.87 NASA Ames USA 2004 10160
3 Altix, Infiniband

5 Dell Thunderbird 38.27 Sandia USA 2005 8000


6 Red Storm
Cray 36.19 Sandia USA 2005 10880
10 Cray XT3
7
NEC Earth-Simulator 35.86 Earth Simulator Center Japan 2002 5120
4
8 MareNostrum Barcelona Supercomputer
IBM 27.91 Spain 2005 4800
5 BladeCenter JS20, Myrinet Center
9 ASTRON
IBM eServer Blue Gene 27.45 Netherlands 2005 12288
6 University Groningen
Jaguar
10 Cray 20.53 Oak Ridge National Lab USA 2005 5200
Cray XT3

26th List / November 2005 www.top500.org page 4

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 450 / 454
Parallel Computing Introduction

Current List (November 2008):

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 451 / 454

Parallel Computing Introduction

Geographic location of top 100 supercomputers:

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 452 / 454
Parallel Computing Introduction

TOP500 List Highlights - November 2008:


Intel dominates the high-end processor market with 75.8 percent of all
systems and 87.5 percent of quad-core based systems.
410 systems are labeled as clusters, making this the most common
architecture in the TOP500 with a stable share of 82 percent.
Quad-core processor based systems have taken over the TOP500 quite
rapidly. Already 336 systems are using them. Seven systems use IBM’s
advanced Sony PlayStation 3 processor with 9 cores.
The average concurrency level in the TOP500 is 6,240 cores per system up
from 4,850 six month ago.
IBM and Hewlett-Packard continue to sell the bulk of systems at all
performance levels of the TOP500. HP took over the lead in systems with
209 systems (41.8 percent) over IBM with 188 systems (37.6 percent).
The U.S. is the leading consumer of HPC systems with 291 of the 500
systems (up from 257). The European share (151 systems – down from 184)
is settling down after having risen for some time, but is still substantially
larger then the Asian share (47 systems – unchanged).

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 453 / 454

Parallel Computing Introduction

The entry level to the list moved up to the 12.64 Tflop/s mark on the
Linpack benchmark, compared to 9.0 Tflop/s six months ago.
The last system on the newest list would have been listed at position 267 in
the previous TOP500 just six months ago.
Total combined performance of all 500 systems has grown to 16.95 Pflop/s,
compared to 11.7 Pflop/s six months ago and 6.97 Pflop/s one year ago.
The entry point for the top 100 increased in six months from 18.8 Tflop/s to
27.37 Tflop/s (which would have been # 9 on the November 2005 list).
The entry level into the TOP50 is at 50.55 Tflop/s (which would have been
# 5 on the November 2005 list).
Of the top 50, 56 percent of systems are installed at research labs and 32
percent at universities. Cray’s XT is the most-used system family with 20
percent, followed by IBMs BlueGene with 16 percent. The average
concurrency level is 30,490 cores per system – up from 24,400 six month ago.
Seven U.S. DOE systems dominate the TOP10.
Roadrunner (# 1) is based on the IBM QS22 blades that are built with
advanced versions of the processor in the Sony PlayStation 3.
The list now includes energy consumption of the supercomputers.

MMAE 517 (Illinois Institute of Technology) Computational Fluid Dynamics


c 2010 K. W. Cassel 454 / 454

You might also like