You are on page 1of 60

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

net/publication/336778117

Orthogonal Collocation Revisited

Article  in  Computer Methods in Applied Mechanics and Engineering · March 2019


DOI: 10.1016/j.cma.2018.10.019

CITATIONS READS
14 436

1 author:

Larry Young
Tilden Technologies, LLC
25 PUBLICATIONS   837 CITATIONS   

SEE PROFILE

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

Orthogonal Collocation Revisited View project

Implementing Methods of Weighted Residuals View project

All content following this page was uploaded by Larry Young on 14 June 2021.

The user has requested enhancement of the downloaded file.


Author prelimiary copy
Computer Methods in Applied Mechanics and Engineering 345, 1033-76 (2019),
https://doi.org/10.1016/j.cma.2018.10.019.

Orthogonal Collocation Revisited


Larry C. Young

Dedicated to Professor Bruce A. Finlayson

Table of Contents
Abstract ...................................................................................................................................2
1. Introduction..........................................................................................................................2
2. Historical Perspective ..........................................................................................................6
2.1 Galerkin Method.............................................................................................................7
2.2 Method of Moments .......................................................................................................7
2.3 Least Squares ................................................................................................................8
2.4 Collocation .....................................................................................................................8
2.5 Other Weight Functions ...............................................................................................11
2.6 Tau Method ..................................................................................................................11
2.7 Boundary Conditions....................................................................................................12
2.8 Finite Elements. ...........................................................................................................13
3. Boundary Value Problem - Catalyst Pellet Problem ..........................................................14
3.1 Orthogonal Collocation Method....................................................................................15
3.1.1 Linear Source, Dirichlet B.C. .....................................................................................16
3.1.2 Linear Source, Third Kind B.C. .................................................................................19
3.2 Method of Moments .....................................................................................................21
3.3 Galerkin Method...........................................................................................................23
3.3.1 Linear Source, Third Kind B.C., Galerkin/Moments ..................................................26
3.4 Mass Conservation and Fluxes ....................................................................................27
4. Parabolic Problem - Falling Liquid Film .............................................................................29
4.1 Continuous Solutions in z ............................................................................................31
4.2 Summary Remarks - Global Methods and Fluxes........................................................37
5. Multidimensional Applications ...........................................................................................38
6. Orthogonal Collocation - Finite Elements ..........................................................................38
6.1 Quadrature in Finite Element Methods ........................................................................40
6.2 C1 Orthogonal Collocation at Gauss Points .................................................................41
6.3 C0 Orthogonal Collocation at Lobatto Points ................................................................42
6.4 Convergence Rate .......................................................................................................45
6.5 Efficiency .....................................................................................................................46
Summary ...............................................................................................................................48
Notation .................................................................................................................................49
References ............................................................................................................................50
Appendix A – Galerkin Method with Flux Boundary Conditions ............................................57
Appendix B – Links to Available Software .............................................................................59

[1]
Abstract
2018 marks the 80th anniversary of Lanczos (1938) and last year was the 50th anniversary of
Villadsen and Stewart (1967), both seminal works in development of the Orthogonal
Collocation Method, also known as the Pseudospectral Method and Differential Quadrature
Method. Due to these milestones, a retrospective on development of the method seems
appropriate. The method’s development as a Method of Weighted Residuals (MWR), with
numerical integration, is reviewed. It is placed into historical perspective and developments in
disparate disciplines are reconciled. We show the equivalence of methods that were thought
by many to be different. Guidance in navigating the minefield of terminology is also provided.

In addition to reviewing the development of the method, we show some refinements which are
not known to most practitioners. Chief amongst these are the correct treatment of boundary
conditions involving fluxes. We demonstrate that a proper treatment of flux boundary
conditions can sometimes reduce errors by several orders of magnitude.

1. Introduction
Orthogonal Collocation (OC), synonymous with the Pseudospectral Method (PS), also goes by
the name of Differential Quadrature Method (DQ). It is a method used for the approximate
solution of differential equations. This method differs from finite difference methods because
the solution is represented as a continuous or piecewise continuous function. For example, if
one is solving for a variable y, the solution is approximated by:
𝑛

𝑦 ≅ 𝑦̃ = ∑ 𝑎𝑘 𝜓𝑘 (𝑥) (1)
𝑘=1
Where x represents one or more spatial coordinates, the a are adjustable parameters which
could be time dependent, and the ψ are called trial functions or basis functions. The trial
functions usually, but not always, satisfy the boundary conditions. The collocation procedure
satisfies the differential equation exactly at a set of collocation points, xk. This criterion
produces values for the adjustable parameters.

By its various names (OC, PS and DQ), the Orthogonal Collocation method has been applied
extensively in such diverse fields as: solid and structural mechanics, fluid mechanics, elasticity,
computational chemistry and chemical physics, quantum mechanics, biomedical modeling,
optimal control, acoustics, geophysics and seismic modeling, heat transfer, mass transfer,
water resource modeling, petroleum reservoir simulation, chemical reaction engineering,
meteorology and atmospheric modeling, and electromagnetism. The number of books and
articles number more than 100,000. For this reason, comprehensive coverage is not feasible.

This paper’s primary focus is on development and formulation of the method rather than
applications. It is further restricted to problems where a trial function representation is applied
to one or more spatial coordinates, such as in a boundary value problem. The application of

[2]
collocation methods to initial value problems is another large area of study [Hairer, et al.
(1993), Hairer and Wanner (1996)]. Only nonperiodic problems in finite domains are
considered, but these other cases are addressed in the many references provided.

We distinguish between several different types of trial functions. In this article, all trial functions
are polynomials. In early applications, the trial functions were often simple monomials, xk. In
later applications the trial functions were orthogonal polynomials, usually Legendre or
Chebyshev polynomials. We call this a modal approximation since the adjustable parameters
are analogous to the modes in a Fourier series. The method is more intuitive and finite-
difference-like if the adjustable parameters represent the approximate solution at the
collocation points rather than polynomial coefficients, i.e. 𝑎𝑘 = 𝑦̃(𝑥𝑘 ). In this case it is called a
nodal approximation and the trial solution is:
𝑛+1

𝑦̃ = ∑ 𝑦̃(𝑥𝑖 )ℓ𝑖 (𝑥) (2)


𝑖=0
where the trial functions, ℓi(x), are Lagrange interpolating polynomials:
𝑛+1
(𝑥 − 𝑥𝑗 )
ℓ𝑖 (𝑥) = ∏
(𝑥𝑖 − 𝑥𝑗 )
𝑗=0
𝑗≠𝑖
The choice of modal or nodal representations is largely a matter of convenience. In the
absence of computer rounding errors all choices produce absolutely identical approximations.
This important fact is obscured in many texts that emphasize the importance of orthogonal
polynomial trial functions.

The easiest way to treat multidimensional problems is by using combinations of the one-
dimensional trial functions. For example, Eq. (2) can be extended by the addition of a second,
z, coordinate by:
𝑛𝑥 +1 𝑛𝑧 +1

𝑦̃ = ∑ ∑ 𝑦̃(𝑥𝑖 , 𝑧𝑗 )ℓ𝑖𝑥 (𝑥) ℓ𝑗𝑧 (𝑧) (3)


𝑖=0 𝑗=0
Trial functions composed of one dimensional trial functions like this are called tensor product
trial functions.

Finally, we distinguish between global trial functions where a single continuous polynomial
approximates the solution and piecewise continuous trial functions called finite element or
spectral element approximations. These trial functions are also called shape functions. In finite
elements, polynomial trial functions are continuous within subdomains called elements but
have limited continuity at the interface between elements. When the elements are distorted
using a polynomial to describe their boundaries, they are called isoparametric elements.

Lanczos (1938,1956) was an early adopter of modal approximations. Villadsen and Stewart
(1967) were the first to use nodal approximations in this context. Neither one considered finite

[3]
element trial functions. In this article, we will concentrate on global approximations, but will
show how the method extends to finite elements in a natural way.

The distinguishing feature of orthogonal collocation is that the collocation points are chosen to
be roots of orthogonal polynomials. The collocation points are usually the roots of Chebyshev
polynomials of the 1st or 2nd kind, or the base points of Gaussian, Radau or Lobatto
quadrature. Gaussian quadrature base points are the roots of Legendre polynomials. All these
choices are the roots of Jacobi polynomials, which possess the orthogonality:
1
(𝛼,𝛽) (𝛼,𝛽)
∫ (1 − 𝑥)𝛼 (1 + 𝑥)𝛽 𝑃𝑛 (𝑥)𝑃𝑚 (𝑥)𝑑𝑥 = 0 𝑓𝑜𝑟 𝑛 ≠ 𝑚 (4)
−1
where α = β = -½, 0, +½, +1 for the Chebyshev 1st kind, Legendre (Gauss), Chebyshev 2nd kind
polynomials, and for those whose roots are Lobatto quadrature base points, respectively. For
increasing n, these polynomials are alternating symmetric and antisymmetric about the
centerline. Radau points are asymmetric with α = 0, β = 1 or α = 1, β = 0. The right half of the
symmetric polynomials along with their extrema are compared in Fig. 1 for n = 8. What should
be noted in Fig. 1 is that as α increases the roots are shifted away from the boundary and
there is a greater variation in the extrema. The endpoint for the Jacobi (1,1) polynomial is off
scale at 9. In all cases the points are more tightly clustered near the boundaries, with spacing
proportional to 1/n2. The 4

Chebyshev polynomials of the 1st 3


Jacobi/Lobatto
kind are the favorite ones for Chebyshev 2 nd
interpolation, since in general 2 Legendre/Gauss
they minimize the maximum error extrema Chebyshev 1 st
Pn

1
[Lanczos (1956) p. 245,
Hildebrand (1987) p. 469]. Note
0
that in Fig. 1 the Chebyshev
polynomials are normalized like -1
extrema
other Jacobi polynomials rather
than the traditional way. -2
0 0.2 0.4 x 0.6 0.8
Fig. 1 Polynomials and extrema for n = 8,  =  = -½, 0, ½, 1
Orthogonal polynomials are closely tied to the theory of accurate numerical integration
methods. For a general reference to orthogonal polynomials and quadrature, the reader is
directed to Hildebrand (1987) and Krylov (1962). The selection of these points is related to
numerical integration of the form:
1 𝑛

∫ 𝑓(𝑥)𝑑𝑥 ≅ 𝑊0 𝑓(0) + ∑ 𝑊𝑖 𝑓(𝑥𝑖 ) + 𝑊𝑛+1 𝑓(1) (5)


0 𝑖=1
Fig. 2 illustrates the point locations and quadrature weights for n = 6. All but Radau points are
symmetric about the centerline. For Gaussian quadrature, endpoint weights are not used, W0 =
Wn+1 = 0, while the Radau-Left quadrature shown has W0 > 0 and Wn+1 = 0. There is a mirror
image Radau-Right formula where the weights and points are switched around. For Lobatto
and Chebyshev (of the 2nd kind) points W0 = Wn+1 > 0. Many claim the approximation of

[4]
boundary conditions is improved by 0.3
nonzero weights at the boundaries;
however, we will show that the exact 0.25

opposite is sometimes true.


0.2

The various quadrature formulas can

w
0.15
calculate exact integrals when the Lobatto
integrand of Eq. (5) is a polynomial of Gauss
0.1
degree 2n+1, 2n, 2n-1 and n+1 for Chebyshev
Radau Left
Lobatto, Radau, Gauss and Chebyshev 0.05
points, respectively. For Chebyshev
points, the Clenshaw-Curtis (1960) 0
0 0.2 0.4 x 0.6 0.8 1
quadrature is formally less accurate,
Fig. 2 Quadrature points and weights
which is certainly not obvious from
examination of Fig. 2. Some claim it is just as accurate in many applications [Trefethen
(2008)]. Clenshaw-Curtis quadrature is distinct from Chebyshev-Gauss quadrature which has
accuracy O(2n) when the radical 1/√1 − 𝑥 2 is included in the integrand of Eq. (5).

With this introduction, we can proceed with the main part of the paper. We begin with Section
2, a historical review of the development of collocation and related methods called Methods of
Weighted Residuals (MWR). Then two linear one dimensional example problems are
considered with global trial functions: a boundary value problem in section 3, the catalyst pellet
problem, and a parabolic problem in section 4, the falling liquid film problem. These problems
are intentionally simple to enable a clear description of the method and some potential pitfalls if
it is not correctly applied. Chief amongst the pitfalls is the correct treatment of flux boundary
conditions. The problems are solved with global collocation, Galerkin and moments (or tau)
methods. Although the examples are simple, the methods presented carry over to nonlinear
multidimensional problems, section 5. In section 6 we show how the global methods extend
naturally to finite elements using the catalyst pellet problem as an example.

One goal of this article is to present a unified and consistent description of methods, and to
show relationships between methods, including the equivalence of methods which are
normally considered to be different. Also, this subject area can be a virtual minefield of
conflicting terminology due to different naming conventions and so forth. To aid the reader, we
have endeavored to include alternate names for various quantities and italicize their definition
when first presented.

Designation of the roots or collocation points is one of the more confusing terminology issues
one encounters. For examples, the names for the roots considered are:
− Chebyshev polynomials of the 1st kind also called: Chebyshev-Gauss or abbreviated
CG points or simply Chebyshev points or a roots grid

[5]
− Chebyshev polynomials of the 2nd kind also called: Chebyshev Extrema or
Chebyshev-Gauss-Lobatto or abbreviated CGL points or simply Gauss-Lobatto or
Chebyshev points
− Legendre polynomials also called: Gaussian quadrature base points or abscissa or
Legendre-Gauss or abbreviated LG points or simply Legendre or Gauss points
− Jacobi Polynomials (α = 0, β = 1 or α = 1, β = 0) also called: Radau points, Radau
quadrature base points or abscissa, Legendre-Gauss-Radau or abbreviated LGR points
− Jacobi Polynomials (α = β = 1) also called: Legendre Extrema points or Jacobi points
or Lobatto quadrature base points or abscissa or Legendre-Gauss-Lobatto abbreviated
LGL points or simply Gauss-Lobatto or Lobatto points

Beware of the terms Chebyshev and Gauss-Lobatto since both are thrown around loosely for
two different sets of points. In this article, we will not consider collocation at the roots of
Chebyshev polynomials of the 1st kind, so Chebyshev with no clarification designates
Chebyshev polynomials of the 2nd kind. The other points in Fig. 2 will be called by the names
found in most standard numerical analysis texts, i.e. Gauss, Lobatto or Radau (left or right)
points [Hildebrand (1987)].

2. Historical Perspective
The collocation method is one of a family of methods collectively called Methods of Weighted
Residuals (MWR) [Finlayson (1972)]. These methods approximate the solution with a linear
combination of trial functions like Eq. (1) or (2). For example, suppose we seek a solution of
the equation:
𝑦𝑥𝑥 + 𝑔(𝑥, 𝑦, 𝑦𝑥 ) = 0 (6)
for 0 < x < 1, with y(0) = y0 and y(1) = y1. The residual is formed by substituting the approximate
solution into the differential equation:
𝑛

𝑅(𝑥, 𝒂) = ∑ 𝑎𝑘 𝜓𝑘𝑥𝑥 + 𝑔(𝑥, 𝑦̃, 𝑦̃𝑥 ) (7)


𝑘=1
If the residual is zero for all x, an exact solution has been achieved. In general, this will not be
possible, so the parameters, a, are selected so the residual will be small in some sense. If the
residual is small for all x, Eq. (1) will be a good approximation to the true solution. MWR
achieves this goal by forcing the function to zero in a weighted average sense:
1
∫ 𝑤𝑘 (𝑥)𝑅(𝑥, 𝒂)𝑑𝑥 = 0 for 𝑘 = 1, … , 𝑛 (8)
0
The wk are called weight or test functions. Different choices for the weight functions lead to
fundamentally different forms of MWR. Finlayson and Scriven (1966) review the early
development of these basic procedures. The MWR were created in the early 20th century to
generalize the Ritz (1908) variational procedure, which has limited applicability. The name
MWR came about from comments by Courant (1924) at the First International Congress of
Applied Mechanics. Crandall (1956, p. 147) was first to illustrate the methods in a unified way

[6]
and following Courant stated the basic principal is that “weighted averages of the residuals”
should vanish. These comments led to the name Methods of Weighted Residuals. They were
called error distribution principals by Collatz (1961) and Ames (1965). The early literature
sometimes called them direct methods. They are also called projection methods,
orthogonalization methods or orthogonal projection methods. The popular name spectral
methods is of unknown origin but seems to have originated about the time of Orzag (1972) and
was initially applied only to global trial functions.

Different choices for the n weight functions produce different methods, we will consider the
following methods:
1. 𝑤𝑘 = 𝜓𝑘 (𝑥), Galerkin method
2. 𝑤𝑘 = 𝑥 (𝑘−1) or 𝑃𝑘 (𝑥), method of moments
𝜕𝑅
3. 𝑤𝑘 = 𝜕𝑎 , least squares
𝑘
4. 𝑤𝑘 = 𝛿(𝑥 − 𝑥𝑘 ), collocation method
There are a few other methods which are not listed above. Several texts include the tau
method of Lanczos (1938, 1956) as a MWR, which is discussed in section 2.6.

2.1 Galerkin Method. The Galerkin method weights the residual by the trial functions. For
problems which can be treated with a variational procedure, the Galerkin method is equivalent.
It was described by Galerkin (1915) in a study of elastic equilibrium and the stability of rods
and plates. An earlier paper by Bubnov (1913) had many similarities, but Galerkin was the first
to describe the method without any reliance on a variational principal. Due to these
relationships it is called the Bubnov-Galerkin method in Russian literature and often the
Raleigh-Ritz-Galerkin method in western literature. The Galerkin method is the most heavily
used MWR and can be considered the gold standard for these methods.

2.2 Method of Moments. The basic ideas behind the method of moments were first described
by N.M. Krylov at a session of the All Ukraine Academy of Sciences in 1926 [Krylov
(1926,1961), Lucka and Lucka (1992)]. The ideas behind the original method are far more
general than weighting by simple monomials. Originally, given some linear differential operator,
L, the weight functions were chosen as 𝑤𝑘 = 𝐿 𝜓𝑘 . The method was further developed by
Kravchuk (1926,1932). Shuleshko (1959) provided additional elaboration. Kravchuk1 spent
much of his short and tragic career studying the various conditions on the linear operator and
trial functions and resulting convergence rates. The generality afforded by this choice of weight
functions includes special cases of the Galerkin and least squares methods, while it also

1 Mykhailo Kravchuk (1892-1942) was the foremost Ukrainian mathematician of the 20th century. He entered the
University at Kyiv in 1910 studying mathematics and physics, teaching after 1917. He endured many hardships
following the Bolshevik revolution and the subsequent loss of Ukranian independence in 1922, but despite the
obstacles, rose to full professor in 1925. In 1938, after years of public service as an educator and academician,
he fell victim to Stalin’s purges and was sent to a Gulag death camp in Siberia. He succumbed to the harsh
treatment in 1942.

[7]
admits eigenfunctions of other linear operators, e.g. Legendre polynomials. Of course, if the
residual is orthogonal to Legendre polynomials it is also orthogonal to all the monomials, x(k-1),
up to the same degree.

The idea of using monomials is apparently due to Yamada (1947) as described by Fujita
(1953) in his study of diffusion in gels. Weighting by monomials can lead to ill conditioned
approximations and some have rejected it for that reason [Boyd (2000) p. 62]. In the western
literature today, the method of moments is generally considered to be weighting by monomials,
but in the current paper we include equivalent methods, such as weighting by Legendre
polynomials.

A basic idea behind the Galerkin and moments methods are that if the weight functions are
members of a complete set, then as n ⟶ ∞ the residual must converge to zero. Of course,
only a finite number of trial functions are used in practice, but the property of completeness
guarantees that the sequence will converge.

2.3 Least Squares. Least squares is a very old idea for approximating functions. Here, it is
used to force the residual to approximate zero. In the context of solving differential equations,
Finlayson and Scriven (1966) attributed its first use to Picone (1928). However, we note the
method was discussed in an earlier paper by Krylov (1926).

The least squares method works well for simple differential equations. However for equations
with nonlinearity and many terms it is more difficult to implement. For this reason, its popularity
has waned as more complex problems have been tackled.

2.4 Collocation. The collocation method dates from the 1930’s. Early work on the method is
due to Slater (1934), Kantorovich (1934), Frazer, et al. (1937) and Lanczos (1938).
Kantorovich called it the interpolation method, Lanczos called it the method of selected points,
while Frazer, et al. applied the name collocation, citing its definition in the Oxford dictionary.

Using all three names we can say the method interpolates the residual to zero at selected
collocation points. In the context of MWR, the weighting functions can be viewed as Dirac delta
functions. Since integration is not required, it is simpler to apply. Simplicity is its most attractive
feature.

Slater applied the method to study energy bands in metal. A translation of the Kantorovich
article is not available, but it is discussed in other articles. Frazer, et al. compared equally
spaced collocation to the Galerkin and least squares methods on several problems. Frazer, et
al. argued that the method would converge whenever the Galerkin or least squares methods
converge. However, Karpilovskaya (1963) points out flaws in their argument. Since collocation
is basically interpolation of the residual, the Runge (1901) phenomenon can occur. A good
example of problems associated with equally spaced collocation can be found in Bert and

[8]
Malik (1996). Frazer, et al. were apparently aware of the Runge phenomenon, but did not
observe it for their problems.

Lanczos used Chebyshev polynomial trial functions and collocated at the roots of Chebyshev
polynomials of the 2nd kind. Clenshaw and Norton (1963) collocated at the same points and
discuss several implementation details for nonlinear problems. Wright (1964) collocated at the
roots of Chebyshev polynomials of the 1st kind, arguing that this choice would minimize the
maximum deviations of the residual from zero. However, a uniform distribution of the residual
does not produce a uniform error in the solution. The use of Chebyshev roots was an important
advance, since the Runge phenomenon does not occur. Villadsen (1970) reviews other early
Chebyshev applications.

Villadsen and Stewart (1967) made a further improvement by selecting the roots of Jacobi
polynomials that are the base points of accurate numerical integration schemes, e.g. Gaussian
or Lobatto quadrature. They chose the name Orthogonal Collocation for their variation on the
method. They explained that collocation at Lobatto points was equivalent to numerical
integration of the Galerkin method. An accurate (sometimes exact) approximation of the
Galerkin method is achieved. They also improved the method by formulating it with nodal
values rather than polynomial coefficients. This modification produced a more intuitive, finite
difference like method and facilitated automation of the method for general applications.
Despite the title of the article, they demonstrated the method for: a boundary value problem,
the parabolic transient catalyst pellet problem, eigenvalues for the Graetz problem and the
elliptic 2-dimensional Poisson equation. All but hyperbolic equations were covered.

In the 1970’s, the development of collocation methods occurred in three threads: (1)
Orthogonal Collocation (OC), (2) Pseudospectral (PS) and (3) Differential Quadrature (DQ).
The OC thread started with the Villadsen and Stewart article, with further development
reported in Villadsen (1970), Finlayson (1972) and Villadsen and Michelsen (1978). These
references describe collocation at Gauss, Lobatto and Radau points. Collocation at Gauss
points was shown to be equivalent to numerical integration of the moments method. In addition
to standard cartesian coordinates, they describe methods for symmetric problems in planar,
cylindrical and spherical coordinates. Nodal differentiation matrices are used exclusively.
These developments led to a flurry of activity applying the method to different problems [see
references in Michelsen and Villadsen (1972, 1981), Finlayson (1974)]. Many of these early
papers demonstrated very favorable comparisons of the method to finite differences. A later
text describing the method is Finlayson (2003).

The Pseudospectral thread began with the work of Orzag (1972), with further developments by
Gottlieb and Orzag (1977). The name pseudospectral is normally synonymous with collocation,
but occasionally it is used to describe other approximations to exact MWR. Orzag applied
collocation to a periodic problem using trigonometric trial functions. He also considered a
nonperiodic linear first order hyperbolic problem with collocation at the roots of Chebyshev

[9]
polynomials of the 2nd kind. For the nonperiodic problem Chebyshev trial functions were used,
and collocation gave an accurate approximation of the Galerkin method. Nodal approximations
were not considered. An important contribution of this work was the use of fast Fourier
transforms (FFT) to perform the calculations. It is not known whether Orzag was aware of the
earlier work of Villadsen and Stewart, but it was not referenced.

FFT was the spark that ignited work in the PS thread. Subsequent work investigated various
methods of solution using FFT. A strong mathematical foundation for the method was laid and
the solution of periodic problems was highly developed. Most of the early work on nonperiodic
problems employed collocation at Chebyshev points with a modal approach using Chebyshev
trial functions. Nodal trial functions and collocation at Lobatto points were adopted later on, but
Villadsen and Stewart are rarely credited for these innovations. Some excellent texts describe
the method, solution algorithms and applications [Canuto, et al. (1988, 2006, 2007), Bernardi
and Maday (1997), Trefethen (2000), Boyd (2000) , Peyret (2002), Shen, et al. (2011)].

The Differential Quadrature Method originated with Bellman and Casti (1971), Bellman, et al.
(1972), and Bellman (1973, p. 244). The first was a short paper advancing the idea of using a
nodal differentiation matrix in the solution of differential equations. The second paper
developed the method further and demonstrated it for both first and second order equations in
one dimension. However, the paper provided few details regarding boundary condition
treatment. Bellman, et al. (1972) developed the differentiation matrices for collocation at Gauss
points. The use of a nodal differentiation matrix was not a new idea. Nielsen (1956) presents
formulas for them with arbitrary nodal locations. Nodal collocation at Gauss points was already
well established in the OC literature. Nevertheless, the method was adopted for many
engineering applications. Bert and Malik (1996) and Bellomo (1997) give excellent reviews of
developments in DQ. The method was initially developed independently of work in the OC and
PS communities. Apparently, it was not recognized as a form of collocation until the late 1980’s
[Quan and Chang, 1989)]. A reference text on the method is Shu (2000).

Ideas have been exchanged between the different threads of development; however, their
source are not always referenced. Chebyshev points have historically been popular in PS
applications and are used in DQ as well. Villadsen and Stewart were the first to advocate
collocation at Gauss and Lobatto points. Lobatto points, also called Legendre-Gauss-Lobatto
or LGL points, have become very popular in the PS literature. Gauss points, also called
Legendre-Gauss or LG points, are popular in OC applications, but not as popular in the PS
and DQ communities. The nodal formulation, pioneered by Villadsen and Stewart, is popular
with “the engineers” due to its similarity to finite differences. Nodal approximations are used in
OC, PS and DQ applications. Symmetric problems in planar, cylindrical and spherical
coordinates are common in OC applications, while problems with periodic solutions are
covered almost exclusively in the PS literature.

[10]
2.5 Other Weight Functions. Some authors take a more general view and refer to virtually all
MWR weightings as Galerkin or variational methods. Alternative weighting schemes are
sometimes called Petrov Galerkin methods. For Chebyshev methods the weights in Eq. (8)
usually include the radical 1/√1 − 𝑥 2 . This term is necessary to exploit the orthogonality of the
polynomials and to achieve accurate integration with Gauss-Chebyshev quadrature. In naming
methods, most authors do not distinguish between methods which include the radical. It would
be more appropriate to add Chebyshev to the name, e.g. Chebyshev-Galerkin method. The
naming of methods with alternative weight functions is yet another area where terminology can
be confusing. In the current article, methods which include the radical will not be used.

2.6 Tau Method. The tau method was originated by Lanczos (1938, 1956) as a method for
approximating functions which can be described by differential equations. The basic idea
behind the method is that rather than find an approximate solution to the differential equation,
why not find an exact solution to an approximate differential equation. For simple problems, he
represented the approximate differential equation as the original one with the addition of a
small perturbation term of the form τPm(x), where τ is a constant and Pm(x) is a high order
orthogonal polynomial, i.e. m ≈ n. Depending on the complexity of the problem, several terms
may be required. These perturbation terms are the residual as defined by Eq. (7), so in general
Lanczos’ idea was to set:
𝑅(𝑥, 𝒂) = 𝜏0 𝑃𝑚 (𝑥) + 𝜏1 𝑃𝑚+1 (𝑥) + ⋯ (9)
where the number of terms required depends on the nature of the differential equation.
Lanczos describes a recursive procedure for determining the values of the τ parameters and
the coefficients in the approximation, i.e. a in Eq. (1).

The tau method was further developed in a series of articles by Ortiz and coworkers [e.g. Ortiz
(1975)]. Several solution methods are described which do not involve integration or
orthogonalization as required by most MWR. Later papers [El-Daou, et al. (1993), El-Daou and
Ortiz (1997)] explore the equivalence between the tau method and other methods. The
relationship between the tau method and the integrated MWR is obvious. Since most MWR
make the residual orthogonal to the first few members of a complete set, the residual will be
proportional to the first neglected term and possibly higher terms. If tau methods are defined to
include all methods with a residual term like Eq. (9), then virtually all MWR (including the
Galerkin method) would be tau methods. This idea seems silly. Instead, we view the tau
method as an alternate solution procedure for MWR which does not involve integration. The
lack of a need for integration is the method’s distinguishing feature, but it also causes its
applicability to be somewhat limited.

The tau method as presented in the spectral literature [e.g. Gottlieb and Orzag (1977, p. 11),
Canuto, et al. (1988, p. 10, 2006, p. 21), Boyd (2000) p. 473] bears little resemblance to the
method described by Lanczos (1956). Unlike the algorithms of Lanczos and Ortiz and
coworkers, the spectral literature describes a method which requires integration. The
Legendre-tau method makes the residual orthogonal to the first few Legendre polynomials, so
[11]
it is equivalent to the method of moments. The Chebyshev-tau method makes the residual
orthogonal to Chebyshev polynomials with the additional weight factor of 1/√1 − 𝑥 2 [Johnson
(1996)], i.e. a Chebyshev-moments method. In the spectral literature, the boundary conditions
are usually satisfied using side conditions, creating a messy matrix structure. Whether the
boundary conditions are satisfied by the trial functions or whether they are imposed by side
conditions makes no difference in the final result. Section 3.2 describes a nodal form of the
method with a symmetric matrix structure.

The tau method offers an interesting view of the residual which we will revisit when discussing
the examples. However, as practiced in the spectral literature it is identical to the method of
moments. In this article, we will use the traditional name method of moments, but the reader
should be aware that the tau method is equivalent. The residual for all the MWR with
polynomial trial functions can be represented by Eq. (9), so the distribution of residual errors
can be shifted by the choice of weighting function, cf. the polynomials in Fig. 1.

2.7 Boundary Conditions. Descriptions of MWR distinguish between three different


approaches: (1) interior methods, (2) boundary methods and (3) mixed methods. An interior
method employs trial functions which obey the boundary conditions, so the parameters are
chosen to approximate the differential equation. In boundary methods, the trial functions obey
the differential equation and the parameters are chosen to approximate the boundary
conditions. In mixed methods, the parameters are adjusted to approximate both the differential
equation and boundary conditions. The foregoing sections have primarily addressed
approximations to the differential equation, which is usually the biggest challenge.

When the boundary conditions are approximated, the residual of the boundary conditions may
be required to meet their own independent weighted residual criteria or they may be enforced
with criteria based on a weighted combination of interior and boundary residuals.

For the moments (or tau) method, all boundary conditions must be satisfied exactly. They may
be satisfied through the selection of the trial functions or by using side conditions [Boyd (2000),
pp 111-115].

For the Galerkin method, the boundary conditions can be divided into two types – essential
boundary conditions and natural boundary conditions [Finlayson (1972)]. For second order
equations, Dirichlet conditions or other conditions on the value of the solution variable are
essential and must be satisfied exactly. Conditions of the 2nd or 3rd type (Neuman or Robin) or
others involving derivatives or fluxes are natural boundary conditions. Natural conditions can
be handled in one of two ways. They can be satisfied exactly, or they can be approximated
using a combined weighting of interior and boundary residuals. The natural treatment is also
called a weak formulation of the boundary conditions. Exact satisfaction of the boundary
condition is called a strong treatment or boundary collocation since, like collocation in the
interior, the residual of the boundary condition is zero.

[12]
Although not a fundamental requirement, collocation implementations normally use a strong
treatment of boundary conditions [e.g. Finlayson (1972), Villadsen and Michelsen (1978), Bert
and Malik (1996), Bellomo (1997), Trefethen (2000), Boyd (2000), Peyret (2002)]. Note that
these references include the OC, PS and DQ threads of development. There have been a few
references that describe an alternative weak formulation of flux boundary conditions [Canuto,
et al. (1988, 2006), Funaro (1992), Shen, et al. (2011)], but these texts also cover the strong
treatment, and none claim a significant benefit to a weak formulation. Consequently, most
applications use a strong treatment or boundary collocation for all boundary conditions.

Although not widely publicized, early applications of OC uncovered a difficulty with flux
boundary conditions [Ferguson (1971), Finlayson (1972), Elnashaie and Cresswell (1973)].
Ferguson (1971) described an integral procedure which circumvented the problem, but its
applicability is limited. It was also suggested that, when the method approximates a Galerkin
method, flux boundary conditions should be formulated as natural/weak boundary conditions
[Young (1977)]. The problem with flux conditions was absent when quadrature weights were
zero at the boundary, i.e. Gauss points, which flies in the face of the commonly held notion that
nonzero boundary weights somehow aid in the approximation of boundary conditions. The
examples in sections 3 and 4, provide a close look at this issue.

2.8 Finite Elements. In the 1970’s there was an explosion of interest in finite element
methods [Zienkiewicz (1971), Strang and Fix (1973), Hughes (1987)]. Its success in structural
mechanics lead to an interest in other application areas. Lanczos and Villadsen and Stewart
considered only global approximations. However, Villadsen and Stewart’s idea of collocating at
quadrature points was soon used to extend OC for use with finite elements. First and most
popular is collocation at Gauss points with early articles by DeBoor and Swartz (1973),
Douglas and Dupont (1973), and Cary and Finlayson (1975). Finite element collocation at
Gauss points is a method with continuous derivatives (C1 continuity) at element interfaces.
Several texts are available which describe the method [Davis (1984), Lapidus and Pinder
(1999), Finlayson (2003)].

Collocation at Lobatto points extends to a finite element method with simple continuity, C0
continuity. The Hybrid-Collocation-Galerkin method was developed first [Diaz (1975), Dunn
and Wheeler (1976), and Wheeler (1977)]. At almost the same time and independently the
Lobatto-Galerkin method was developed [Gray (1977), Young (1977), Hennart (1982)].
Heretofore, these two methods were considered distinctly different, but in section 6 we show
they are equivalent. Leyk (1986,1997) appears to have developed the same method. Later, the
method was developed again by Maday and Patera (1989) and popularized as the hp Spectral
Element method. Texts and monographs describing the method are Vosse and Minev (1996),
Canuto, et al. (2007), Karniadakis and Sherwood (2013). These references credit Maday and
Patera with discovery of the method even though their work came several years after its initial
development.

[13]
Since the discussion of finite element collocation methods gets into the large area of reduced
quadrature (i.e. less than exact integration) in finite element methods, further discussion is
deferred until section 6, after the background information is presented in sections 3 and 4.

3. Boundary Value Problem - Catalyst Pellet Problem


Consider reaction and diffusion in a porous catalytic slab. This differential equation is one of
the most ubiquitous ones in engineering and physics. It is analogous to heat conduction in a
slab with a source dependent on temperature and position. If the source, r, is constant, the
equations describe laminar flow between parallel plates. It also describes the position of an
elastic string, where r characterizes the load. The governing equations are:
𝑑2𝑦
+ 𝑟(𝑥, 𝑦) = 0 (10)
𝑑𝑥 2
where 𝑟(𝑥, 𝑦) = 4𝜑 2 𝑟̂ (𝑥, 𝑦) and y is the fractional conversion, 0 for no reaction and 1 for
complete reaction. φ is called the Thiele modulus. The rate dependence on x permits a
reactivity that depends on position. φ is defined such that the average value of r̂ (x,0) is 1. The
boundary conditions are either first kind, Dirichlet, boundary conditions:
𝑦(0) = 𝑦(1) = 0 (10a)
or third kind, Robin, boundary conditions:
𝑑𝑦 𝑑𝑦
| = 2𝐵𝑖0 𝑦(0) and − | = 2𝐵𝑖1 𝑦(1) (10b)
𝑑𝑥 𝑥=0 𝑑𝑥 𝑥=1
In Eq. (10b), the Biot numbers, Bi, account for an external transfer resistance. As Bi tends to
infinity, Eq. (10b) reduces to Eq. (10a). This problem is often symmetric about the centerline,
but here we allow for a nonsymmetric profile. Symmetric problems in various geometries can
be efficiently treated using polynomials in x2. The Thiele modulus and Biot numbers are
defined using half the thickness of the slab, so the factors of 4 and 2 are required when the full
slab is considered.

For a simple kth order reaction with the reactivity independent of position, the source term is:
𝑟̂ (𝑥, 𝑦) = (1 − 𝑦)𝑘 (11)

For a first order reaction, k = 1, and a single value for Bi, the analytical solution is:
cosh [𝜑(2𝑥 − 1)]
𝑦=1− (12)
cosh(𝜑) + (𝜑/𝐵𝑖)sinh(𝜑)
The quantity of interest from the solution is called the effectiveness factor, η, which gives the
overall rate of reaction relative to that with no diffusional resistance:
1 1
∫0 𝑟(𝑥, 𝑦)𝑑𝑥
𝜂= 1 = ∫ 𝑟̂ (𝑥, 𝑦)𝑑𝑥 (13)
∫0 𝑟(𝑥, 0)𝑑𝑥 0

[14]
The effectiveness factor is basically a normalized boundary flux, since the divergence theorem
in one dimension gives:
𝑑𝑦 1 1
− | = 4𝜑 ∫ 𝑟̂ (𝑥, 𝑦)𝑑𝑥 = 4𝜑 2 𝜂
2 (14)
𝑑𝑥 0 0
Eq. (14) is given the generic name average energy equation, since in a heat transfer setting it
is an overall energy balance and the right-hand side is the average energy generated. Using
the analytical solution, Eq. (12), this normalized flux is:
1
𝜂= (15)
φ[coth(𝜑) + 𝜑/𝐵𝑖]
η is approximately one for φ < ½ and becomes asymptotic when φ > 2. The asymptotic state
corresponds to a condition where all the reactants are consumed near the boundary and the
conversion, y, approaches one at the center.

3.1 Orthogonal Collocation Method


Early developments of the Orthogonal Collocation method state that the trial functions are
orthogonal polynomials, but then use monomials and through transformations formulate
problems in terms of nodal values. Although the approximation turns out the same, we prefer
to dispense with the transformations and develop the method directly. To solve the problem
with a nodal formulation, we start with the trial solution Eq. (2).

The interpolation points in Eq. (2) include the endpoints, x0 = 0 and xn+1 = 1, while the interior
points are the roots of an orthogonal polynomial. Most authors seem almost paranoid about
these boundary points when used with Gauss points in the interior, because the associated
quadrature weights are zero, see Eq. (5) and Fig. 2. They state that nonzero weights are
needed on the boundary to help meet the boundary conditions. There is never any justification
for these statements and, in fact, we will show that in some instances nonzero boundary
weights are a detriment rather than an asset for meeting the boundary conditions.

The residual, Eq. (7), is formed by substitution of the approximate solution into the equation:
𝑛+1 𝑛+1
𝑑 2 ℓ𝑖
∑ 𝑦(𝑥𝑖 ) + 𝑟 (𝑥, ∑ ℓ𝑖 (𝑥)𝑦(𝑥𝑖 )) = 𝑅(𝑥, 𝒚) (16)
𝑑𝑥 2
𝑖=0 𝑖=0
Where y is the vector of nodal values y(xi). With the collocation method, the residual is set to
zero at the interior collocation/interpolation points, xj, j = 1,…,n. Since ℓi(xj) = δij (the dirac delta
function), the resulting equation simplifies to:
𝑛+1
𝑑 2 ℓ𝑖
∑ 𝑦(𝑥𝑖 ) | + 𝑟 (𝑥𝑗 , 𝑦(𝑥𝑗 )) = 0 (17)
𝑑𝑥 2 𝑥
𝑖=0 𝑗

By defining B as indicated below and letting 𝑦𝑖 = 𝑦(𝑥𝑖 ) the equation simplifies to:
𝑛+1

∑ 𝐵𝑗𝑖 𝑦𝑖 + 𝑟(𝑥𝑗 , 𝑦𝑗 ) = 0 (18)


𝑖=0

[15]
The boundary conditions provide two additional conditions, either first kind, Dirichlet, boundary
conditions:
𝑦0 = 𝑦𝑛+1 = 0
or third kind, Robin, boundary conditions:
𝑛+1 𝑛+1

∑ 𝐴0,𝑖 𝑦𝑖 = 2𝐵𝑖0 𝑦0 and − ∑ 𝐴𝑛+1,𝑖 𝑦𝑖 = 2𝐵𝑖1 𝑦𝑛+1 (19)


𝑖=0 𝑖=0
where:
𝑛+1
𝑑ℓ𝑖 𝑑 2 ℓ𝑖
𝐴𝑗𝑖 = | and 𝐵𝑗𝑖 = | = ∑ 𝐴𝑗𝑘 𝐴𝑘𝑖 (20)
𝑑𝑥 𝑥𝑗 𝑑𝑥 2 𝑥
𝑗 𝑘=0
For Dirichlet conditions, Eq. (10a), the boundary values (whether zero or finite) are simply
substituted into Eq. (18). For the third kind conditions, Eq. (10b), most texts and articles
recommend that the boundary conditions be satisfied exactly as in Eq. (19), i.e. boundary
collocation or a strong treatment [Finlayson (1972), p. 101; Villadsen and Michelsen (1978), p.
137; Bert and Malik (1996); Bellomo (1997); Trefethen (2000), p. 137; Boyd (2000), p. 111,
Peyret (2002), p. 59]. Below, boundary derivatives or fluxes are discussed further and a
method superior to Eq. (19) is demonstrated.

To solve the problem, we need only the collocation points, i.e. the roots of the orthogonal
polynomial, x, the derivatives of the Lagrange interpolating polynomials, A and B, and the
quadrature weights, W, for approximating integrals, e.g. Eq. (13). A and B are called
differentiation matrices. The original methods for computing these quantities was crude, but
improvements were quickly developed [Michelsen and Villadsen (1972)]. The roots can be
calculated iteratively, while the weights and differentiation matrices can be calculated from the
polynomial derivatives evaluated at the roots. Alternatively, the roots and quadrature weights
can be calculated from the eigenvalues and eigenvectors of a symmetric tridiagonal matrix
formed from the polynomial recurrence relationships [Golub and Welch (1969)]. Computation
of the differentiation matrices with arbitrary point locations is described by Nielsen (1956).
Many of the texts cited discuss procedures
1
for performing these calculations [e.g.
Villadsen and Michelsen (1978)]. For
software availability in a variety of 0.8

computer languages see Appendix B.


0.6
n=4, Lobatto
3.1.1 Linear Source, Dirichlet B.C.
y

n=4, Gauss
Fig. 3 shows solutions of Eqs. (10) and n=4, Chebyshev
0.4
Exact
(10a) for a first order reaction and φ = 5,
i.e. Eq. (11) with k = 1. Approximate 0.2
solutions are shown with n = 4 for
collocation at Lobatto, Gauss and 0
0 0.2 0.4 0.6 0.8 1
Chebyshev points. With this relatively x
Fig. 3 Solutions of Eqs. (10) and (10a), n = 4,  = 5

[16]
high reaction rate most of the reaction occurs near the boundary. The fifth order polynomial
can only approximate the sharp profile by oscillating about the exact solution. Since this
problem is symmetric about x = 0.5, the coefficient of the fifth order term is zero.
12
Table 1
Coefficients of Residuals 10 Lobatto n = 6
Chebyshev n = 6
Gauss Chebyshev Lobatto Gauss n = 6
n (0.0) ( ) (1,1)
8
𝑃𝑛 (𝑥) 𝑃𝑛 0.5,0.5 𝑃𝑛
6
4 33.6927 16.4722 9.0058

Residual
6 5.6737 2.8395 1.5242 4
8 0.5711 0.2892 0.1544
2

0
Fig. 4 shows the residual for the same
problem and n = 6. The nature of the -2

residual was discussed briefly in section -4


0 0.2 0.4 0.6 0.8 1
2.6. For even n, each of the residuals is an x
Fig. 4 Residual for Eq. (10) and (10a), n = 6,  = 5
nth order Jacobi polynomial with α = β = 0,
+½, +1 for Gauss, Chebyshev and Lobatto points, respectively. The residuals follow Eq. (9),
but each choice of points makes the residuals orthogonal to a different Jacobi polynomial, see
Fig. 1, so the residual errors are distributed differently. The values of the proportionality
constant, τ, given in Table 1 decrease at a 0
10
rapid rate as n is increased.
-2
10
Fig. 5 shows the L2 error norms versus n for
-4
10
three choices of points. A plot of the L1 error
L2 Error

norm looks very similar. All methods show -6


10
the typical exponential convergence with
Lobatto points giving slightly better results. 10
-8

Lobatto
The error with Chebyshev and Gauss points Gauss
-10
10 Chebyshev
averages 1.3 and 1.9 times that with Lobatto
points, which is relatively small since the -12
10
error drops by almost an order of magnitude 2 4 6 8 10 12 14
n
Fig. 5 L2 Error Norm Eqs. (10) and (10a),  = 5
with each increment of n. The largest
residual values occur at the boundary with Lobatto points, but, nevertheless, this choice
produces the smallest errors in the solution.

Using Eq. (15), the normalized boundary flux, η = 0.199983. Numerical quadrature can be
used to calculate this value from the numerical solutions using Eq. (13):
𝑛+1

𝜂 = ∑ 𝑊𝑖  𝑟̂ (𝑥𝑖 , 𝑦𝑖 ) (21)


𝑖=0

[17]
Using Eq. (21) for the cases in Fig. 3, the calculated fluxes are in error by 0.8, -3.1 and -4.3
percent for Lobatto, Chebyshev and Gauss points, respectively. The same quantity can be
calculated using Eq. (14), with the derivatives approximated by:
𝑛+1 𝑛+1
𝑑𝑦 𝑑𝑦
| = ∑ 𝐴0𝑖  𝑦𝑖 and | = ∑ 𝐴𝑛+1,𝑖  𝑦𝑖 (22)
𝑑𝑥 𝑥=0 𝑑𝑥 𝑥=1
𝑖=0 𝑖=0
Using Eq. (22), the derivatives of the solution at the boundaries are in error by -14.3, -10.2 and
-4.3 percent for Lobatto, Chebyshev and Gauss points. These errors are much larger and the
relative accuracy of the methods is a complete reversal of that found with Eq. (21). Only
Gauss points give the same result with either method of flux calculation. Shortly, we will
explain why this is so. For the other two methods, integration gives a far more accurate result.
0
Fig. 6 shows the flux errors for increasing n. 10

Relative to Fig. 5, this graph shows a much 10


-2

greater difference between the calculations. -4


10
For n < 4, the error with all three methods is
relatively large, while for n > 14, some of the Flux Error
-6
10

results are affected by rounding errors. 10


-8

Engineering accuracy is obtained with 4 to 8 Lobatto


points depending on the method and 10-10 Gauss
Chebyshev
accuracy required. In this range, the errors 10
-12 Lobatto Der.
Chebyshev Der.
with Gauss and Chebyshev points are similar,
10-14
while Lobatto points give somewhat greater 2 4 6 8
n
10 12 14
accuracy. Fig. 6 Flux Error, Dirichlet  = 5, k = 1

The understanding of convergence properties of the method has improved greatly since it was
first developed. Ferguson and Finlayson (1972) presented error bounds for Eqs. (10) and
(10a), while Michelsen and Villadsen (1981) give an error expression for the linear problem, k
= 1. Examples of other early work on convergence analysis are in Gottlieb and Orzag (1977)
and Canuto and Quarteroni (1981). More recent work is summarized in many of the reference
books cited in section 2.4 [e.g. Canuto, et al. (2006)].

Since Figs. 5 and 6 are so different let us take a closer look at these measures of the error.
Given the exact solution, y*, the Lp error norm is a measure of the error in the internal profile:
1
1 𝑝
𝜖𝑝 = [∫ |𝑦 − 𝑦 ∗ |𝑝 𝑑𝑥] (23)
0
The error in the normalized flux for this linear source function is:
1
𝜖𝜂 = |∫ (𝑦 − 𝑦 ∗ )𝑑𝑥| (24)
0

These two error measures look similar, especially when p = 1. However, upon closer
examination, it is clearly possible to achieve an exact flux with an imperfect solution which

[18]
oscillates about the exact solution but in a way that gives the correct solution on average. If
one plots y – y* it is obvious that the error at the collocation points is smaller than the average
Lp error. The combination of greater accuracy at the points and accurate quadrature makes for
the accurate flux calculations in Fig. 6.

Villadsen and Michelsen (1978) (p. 85) also note that other methods may produce similar
profile errors, such as shown in Fig. 5, but the moments and Galerkin methods do an
especially good job of balancing the error, so that the flux calculation is very accurate. As we
shall see, collocation at Gauss and Lobatto points have equivalence to moments and Galerkin
methods, respectively. The only study which seems to address this better than expected
performance of certain global approximations is the one by Lanczos (1973).

It is likely the more rapid convergence rate for fluxes with Lobatto and Gauss points is related
to the long known superconvergence property of finite element methods [see Křížek and
Neittaanmäki (1998)]. The superconvergence phenomenon is one where the solution and/or its
derivative converge faster at certain points than the solution on average. Superconvergence
for orthogonal collocation finite element procedures is discussed in Section 6. Since the global
methods considered here are but a single element of a finite element procedure, it is likely the
exceptional accuracy observed here in Fig. 6 is a related phenomenon.

The error curves in Figs. 5 and 6 converge at a supergeometric rate, i.e. with (n)log(n), but are
plotted versus n as is customary. Lobatto points give the best convergence rate. Gauss points
converge at the same rate, but the errors are about 10 to 15 times larger, equivalent to about
one increment of n. The convergence rate with Chebyshev points is roughly half that with
Gauss or Lobatto points. If derivatives are used to calculate the flux, Eq. (22), the results are
poor with Lobatto or Chebyshev points. Gauss points produce the same result regardless of
calculation method. Since all the methods give exponential convergence and the L1 and L2
error norms are similar, one could easily be misled by an incomplete comparison. Significant
differences show up only when fluxes are compared. One of the advantages of orthogonal
collocation or pseudospectral methods is that virtually exact solutions are feasible. The method
with Lobatto points is superior for achieving high accuracy for this example, but only if fluxes
are accurately calculated.

3.1.2 Linear Source, Third Kind B.C.


Now, consider the same problem as above, but with the third kind boundary conditions, Eq.
(10b). It is clear from Eqs. (12) and (14) that φ/Bi gives the relative importance of internal and
external resistance. The shape of the profile is not changed, but the boundary value is scaled
up to account for the external resistance. We first present calculations with φ = 5 as above and
Bi = 10. For these conditions both are important, but the external resistance is less important
than the internal resistance. Later, other values of Bi are considered to determine its effect on
the results.

[19]
Most texts recommend that boundary collocation, Eq. (19), or an equivalent strong method be
used to approximate the boundary condition. Given that Fig. 6 shows poor accuracy of
derivatives (Lobatto and Chebyshev) for calculating fluxes, one might question the suitability of
this approach for all but Gauss points.

Fig. 7 compares the profiles for solutions 1


calculated with n = 4. Note that the boundary
value of y is about 0.3, indicating roughly 30%
0.8
of the resistance is external. The relative
accuracy of the methods appear similar to
that shown in Fig. 3 with Dirichlet conditions. 0.6

y
Lobatto, n = 4
Fig. 8 shows the L2 error norms are almost Gauss, n = 4
the same for Chebyshev and Gauss points Chebyshev, n = 4
0.4 Exact
and averages about 50 percent greater for
Lobatto points. If only the norms are
compared, one might conclude that all 0.2
0 0.2 0.4 x 0.6 0.8 1
methods give similar accuracy. However, Fig.
Fig. 7 Solutions Eqs. (10) and (10b), Bi = 10,  = 5
9 shows a significant disparity in fluxes
calculated with Eq. (21). Since Eq. (12) shows that the shape of the profile is the same, the
greater error must be due to a problem in the formulation. Although, the differences are not
large for n < 6, the differences in convergence rate are almost a factor of 2. As discussed
below, similar flux errors persist even when Bi is so large that a Dirichlet condition is
approached.
0
100 10

-2 10-2
10

-4
10
-4
Flux Error

10
L2 Error

-6
10
-6
10
-8
10

-8 Gauss Gauss
10
Lobatto Coll. 10
-10
Lobatto Coll.
Chebyshev Coll. Cheb. Coll.
-10
10
2 4 6 8 10 12 14 10-12
n 2 4 6 8 10 12 14
n
Fig. 8 L2 Error Norms Eqs. (10) and (10b), Bi = 10  = 5 Fig. 9 Flux Error Eqs. (10) and (10b), Bi = 10  = 5

Although, significant errors occur only in the fluxes for this problem, other equations show
larger errors in the internal profiles as well [e.g Young (2018)].

This problem with flux boundary conditions was pointed out in the early 1970’s [Ferguson
(1971), Elnashaie and Cresswell (1973)]. The problem is clearly evident for two different
problems in Finlayson (1972) (see Table 5.7 and Fig. 5.7). The problem is only briefly

[20]
discussed by Villadsen and Michelsen (1978, p. 248). Other articles attributed the problem to
the strong treatment of flux boundary conditions (see Appendix A) and recommended
collocation at Gauss points for these problems [Young and Finlayson (1976), Michelsen and
Villadsen (1981)]. Unfortunately, many are apparently unaware of the problem, since later texts
continued to recommend Eq. (19) or its equivalent for Chebyshev and Lobatto points. This
problem caused collocation at Lobatto points to fall from favor in the OC community. A superior
alternative to the strong treatment, Eq. (19), can be found by examining other Methods of
Weighted Residuals (MWR) and the foundation of orthogonal collocation.

3.2 Method of Moments


As discussed in section 2.6, the method of moments is equivalent to the tau method as
practiced in the spectral literature [Boyd (2000), Canuto, et al. (2006)]. To solve Eq. (10) by the
method of moments, the residual is weighted by monomials xk for k = 0,…,n - 1; however, as
discussed in section 2.2, weighting by any linearly independent set of n polynomials through
degree n - 1 will give identical results. For example, we could use the first n Legendre
polynomials and the results would be identical except possibly for rounding errors. Another
suitable set of linearly independent polynomials are the Lagrange interpolating polynomials for
only the n interior points. These polynomials are related to those in Eq. (2) by:
𝑥𝑖 (1 − 𝑥𝑖 )
ℓ∗𝑖 (𝑥) = ℓ𝑖 (𝑥) (25)
𝑥(1 − 𝑥)
where the asterisk indicates the reduced polynomial. Weighting the residual function, Eq.(16),
by these polynomials and integrating numerically gives the following problem:
𝑚 𝑛+1 𝑛+1
𝑑 2 ℓ𝑖
∑ [(∑ 𝑦𝑖 | ) + 𝑟 (𝑥𝑘 , ∑ ℓ𝑖 (𝑥𝑘 )𝑦𝑖 )] 𝑊𝑘 ℓ𝑗∗ (𝑥𝑘 ) = 0 (26)
𝑑𝑥 2 𝑥
𝑘=1 𝑖=0 𝑘 𝑖=0

for j = 1,…,n, where xk and Wk designate quadrature base points and weights, respectively,
which are yet to be specified. The method of moments requires that all boundary conditions be
satisfied exactly. For Dirichlet conditions, Eq. (10a), the boundary values are substituted. For
third kind conditions, Eq. (10b), Eq. (19) provides the two extra equations.

For m > n the quadrature base points are naturally different from the nodal interpolation points
used to define the trial functions, Eq. (2). However, if m = n, and the interpolation points
coincide with the quadrature points, some wonderful simplifications occur. For this case, Eq.
(26) simplifies to:
𝑛+1

∑ 𝑊𝑗 𝐵𝑗𝑖 𝑦𝑖 + 𝑊𝑗 𝑟(𝑥𝑗 , 𝑦𝑗 ) = 0 (27)


𝑖=0
Eq. (27) is equal to Eq.(18) when each row is multiplied by the quadrature weight, Wj, so the
equations are equivalent. To make this approach work, we need a quadrature formula that will
give an exact or accurate approximate integration of Eq. (26) with m = n and 0 < xi < 1. The
most accurate quadrature of this type is Gaussian quadrature.

[21]
Let us now examine the accuracy of Gaussian quadrature for integration of Eq. (26). The trial
functions, ℓi(x), are polynomials of degree n + 1, so the second derivative is of degree n – 1.
The weight functions, ℓj*(x) are of degree n – 1. Combining the two terms, the diffusion term is
of degree 2n – 2. The first order reaction term is of degree 2n. Since Gaussian quadrature is
exact for polynomials through degree 2n – 1, integration of the diffusion terms is exact, while
the source term misses exact integration by one degree. So, Eq. (27) evaluated with Gaussian
quadrature is both an accurate approximation of the moments method and equivalent to
collocation at Gauss points, Eq.(18).

Lobatto and Radau quadrature with m = n have sufficient accuracy to integrate all the terms in
Eq. (26) exactly for a first order reaction. However, neither one reduces to a collocation
method because end point terms would appear in Eq. (27). These terms are zero with Gauss
points, because the quadrature weights are zero at both boundaries. This is a case where
quadrature weights on the boundary are not an asset as many claim. Orthogonal collocation at
Lobatto or Radau quadrature base points bear no direct relationship to the moments method.
With Chebyshev points, the associated Clenshaw-Curtis quadrature also has nonzero
quadrature weights at the endpoints. In addition, Clenshaw-Curtis quadrature is not accurate
enough to produce a good approximation to the moments method. For these reasons, it also
bears no direct relationship to the moments method.

To achieve an exact representation of the moments method with a first order reaction,
𝑟̂ (𝑥, 𝑦) = 1 − 𝑦, the following integration must be performed more accurately:
1
𝐷𝑗𝑖 = ∫ ℓ𝑗∗ (𝑥)ℓ𝑖 (𝑥) 𝑑𝑥 ≈ 𝛿𝑗𝑖 𝑊𝑗 (28)
0
We call D the mass matrix and define D0,i = Dn+1,i = 0. With Gauss points it is approximated by
the diagonal matrix of quadrature weights shown at the far right. In finite element methods, the
reduction of the mass matrix to a diagonal one is called lumping. Lumping is achieved
automatically here due to the approximate integration. D is also called a capacity matrix, since
it is often associated with time derivative terms.

To list the complete set of equations, it is convenient to combine the boundary and interior
equations by defining:
𝐶𝑗𝑖 = 𝛿𝑗,𝑛+1 𝐴𝑛+1,𝑖 − 𝛿𝑗,0 𝐴0,𝑖 − 𝑊𝑗 𝐵𝑗𝑖 (29)
With this definition, the complete set of equations for the first order reaction with constant
coefficients and third kind boundary conditions is:
𝑛+1 𝑛+1

𝛿𝑗,0 2𝐵𝑖0 𝑦0 + 𝛿𝑗,𝑛+1 2𝐵𝑖1 𝑦𝑛+1 + ∑(𝐶𝑗𝑖 + 4𝜑 𝐷𝑗𝑖 )𝑦𝑖 = 4𝜑 ∑ 𝐷𝑗𝑖 = 4𝜑 2 𝑊𝑗


2 2 (30)
𝑖=0 𝑖=0
C is a symmetric matrix which we call the stiffness matrix. For the full moments method, the
complete matrix problem is not symmetric because of D. However, with the diagonal

[22]
approximation of D (far right of Eq. (28)), the equations are symmetric and positive definite and
equivalent to collocation. One deficiency of orthogonal collocation, pseudospectral or
differential quadrature methods is that self adjoint operators do not lead to symmetric matrix
problems. This development shows that by a simple rescaling of the equations this desirable
feature is achieved with Gauss points. A symmetric matrix problem cuts the calculations for
solution almost in half [Faddeeva, 1959]. Symmetric matrices also have theoretical and
computational benefits for other problems, e.g. the eigenvalue problem in section 4.

3.3 Galerkin Method


To solve the problem with the Galerkin method, the residual is weighted by the trial functions
ℓj(x). Since the Robin boundary conditions reduce to Dirichlet conditions for large Bi, we will
consider only the more general conditions, Eq. (10b). As discussed in section 2.7, the Galerkin
method permits flux boundary conditions to be treated strongly or weakly [Finlayson (1972)].
The weak or natural treatment is most commonly used. However, the normal procedure with
the orthogonal collocation, pseudospectral or differential quadrature method is to satisfy the
boundary conditions exactly, i.e. boundary collocation, Eq. (19). The short development in
Appendix A shows the difficulties with this approach, especially for nonzero quadrature weights
at the boundary. For this reason, we will use the natural boundary condition treatment here.
The natural treatment allows the polynomial greater flexibility to adapt to the solution, while the
boundary condition is satisfied approximately along with the rest of the differential equation.

To solve the problem with the Galerkin method, the residual, Eq. (2), is weighted by the trial
functions ℓj(x), for j = 0,…,n + 1. The equations are converted to the weak formulation by
integrating the second derivative term by parts:
𝑛+1 𝑛+1
𝑑ℓ𝑖 1 1 𝑑ℓ𝑗 𝑑ℓ𝑖
∑ ℓ𝑗 | 𝑦𝑖 − ∫ (∑ 𝑦 − ℓ𝑗 𝑟(𝑥, 𝑦)) 𝑑𝑥 = 0 (31)
𝑑𝑥 0 0 𝑑𝑥 𝑑𝑥 𝑖
𝑖=0 𝑖=0
The first term contains the two boundary derivatives, so we substitute the boundary conditions
and integrate the other terms using a suitable quadrature formula:
𝑚 𝑛+1
𝑑ℓ𝑗 𝑑ℓ𝑖
𝛿𝑗,0 2𝐵𝑖0 𝑦0 + 𝛿𝑗,𝑛+1 2𝐵𝑖1 𝑦𝑛+1 + ∑ 𝑊𝑘 (∑ | | 𝑦 − ℓ𝑗 (𝑥𝑘 ) 𝑟(𝑥𝑘 , 𝑦(𝑥𝑘 ))) = 0 (32)
𝑑𝑥 𝑥 𝑑𝑥 𝑥𝑘 𝑖
𝑘=1 𝑖=0 𝑘

The xk in Eq. (32) designate the quadrature base points, which differ from the nodal
interpolation points for m > n.

Consider quadrature with n interior points. Since the trial functions are polynomials of degree
n+1, the diffusion term is of degree 2n. For a first order reaction, the source term is of degree
2n+2. An n point Gaussian quadrature is exact through degree 2n-1, so neither term would
be integrated exactly. On the other hand, Lobatto quadrature with n interior points gives exact
integration through degree 2n+1, so it gives exact integration for the diffusion term, but misses
exact integration of a linear source term by one degree. This is the same level of discrepancy
found when the moments method is approximated with Gaussian quadrature. Radau
[23]
quadrature is one degree less accurate than Lobatto quadrature, so Radau quadrature also
integrates the diffusion term exactly.

Using Lobatto quadrature with n interior points, Eq.(32) reduces to:


𝑛+1

𝛿𝑗,0 2𝐵𝑖0 𝑦0 + 𝛿𝑗,𝑛+1 2𝐵𝑖1 𝑦𝑛+1 + ∑ 𝐶𝑗𝑖 𝑦𝑖 − 𝑊𝑗  𝑟(𝑥𝑗 , 𝑦𝑗 ) = 0 (33)


𝑖=0
where:
𝑛+1

𝐶𝑗𝑖 = ∑ 𝑊𝑘 𝐴𝑘𝑗 𝐴𝑘𝑖 = 𝛿𝑗,𝑛+1 𝐴𝑛+1,𝑖 − 𝛿𝑗,0 𝐴0,𝑖 − 𝑊𝑗 𝐵𝑗𝑖 (34)


𝑘=0
Since Lobatto quadrature can perform the integration by parts exactly, it follows that the
stiffness matrix, C, can be calculated by either expression above. The right expression is
identical to Eq. (29). Given this relationship, it is clear that at the interior points, Eq. (33) is
identical to Eq. (27) and equivalent to Eq. (18), i.e. collocation.

To examine the different treatment of the boundary conditions, we compare Eqs. (19) and (33)
at only one boundary, since the approximation is the same at the other boundary. Using the
right expression of Eq. (34) for C, the boundary equation at j = 0 is:
𝑛+1 𝑛+1

2𝐵𝑖0 𝑦0 − ∑ 𝐴0𝑖 𝑦𝑖 − 𝑊0 (∑ 𝐵0𝑖 𝑦𝑖 + 𝑟(0, 𝑦0 )) = 0 (35)


𝑖=0 𝑖=0
Eq. (35) differs from (19) by the extra term on the right, which is the boundary quadrature
weight multiplying the residual at the boundary, R(0,y) from Eq. (16). This procedure sets a
weighted average of the two residuals to zero. It is still a form of collocation, since residuals
are set to zero at the boundary. The boundary condition is not satisfied exactly, but the
procedure drives both residuals to zero at an exponential rate.

Eq. (35) is equally valid for Gauss points, since the quadrature weights are zero at the
boundaries and it reduces to boundary collocation. The left expression of Eq. (34) is not valid
for Gauss or Chebyshev points, because the quadrature is not accurate enough to perform the
integration by parts exactly. For an infinite Bi number, Eq. (35) reduces to the Dirichlet
condition, y0 = 0.

For a first order reaction, the full Galerkin method requires a more accurate calculation of the
source term in Eq. (32). The term gives a mass matrix, which is identical to Eq. (28) with the
substitution of the trial functions, ℓ, for the reduced weighting functions in Eq. (32). A more
accurate integration of the source term in Eq. (32), produces an equation identical in form to
Eq. (30). For the Galerkin method, both the stiffness and mass matrices are symmetric. The
mass matrix is full for the Galerkin method and is diagonal or lumped for collocation at Lobatto
points. For more complex rate expressions, a full mass matrix adds calculations which do not
normally improve the accuracy enough to warrant the extra complexity.

[24]
In summary, collocation at Lobatto quadrature base points is an accurate approximation of the
Galerkin method, but only when a natural or weak boundary condition treatment is used.
Collocation at Gauss points is an accurate approximation of the moments method. Collocation
at Radau points is in between and reduces to boundary collocation at one boundary and a
natural treatment at the other.

The Clenshaw-Curtis quadrature is not accurate enough to give a good approximation to either
the Galerkin or moments methods. Chebyshev points approximates MWR with the extra
radical, 1/√1 − 𝑥 2 , in the weight function. The Chebyshev points and quadrature weights are
intermediate between Gauss and Lobatto points and weights (see Fig. 2), so can also be
justified on that basis. For Chebyshev points, we propose the use of a natural boundary
condition treatment, i.e. Eq. (33), with C calculated by Eq. (29) or the right expression of Eq.
(34). We will call this a natural or weak boundary condition treatment also even though the
quadrature does not produce an accurate approximation of Eq. (32).

Formulations using the stiffness matrix and a natural treatment of flux boundary conditions are
referred to as weak formulations. Unlike with Gauss and Lobatto points, the stiffness matrix for
Chebyshev points is not symmetric (except for n < 4). Consequently, almost twice the
computations are required to solve matrix problems. Also note the natural boundary condition
treatment causes boundary rate terms to appear in the approximation, Eq. (35). For a
nonlinear rate expression, all n + 2 equations are nonlinear, whereas with Gauss points the
two boundary equations are linear. These linear equations can be eliminated initially so only n
nonlinear equations must be solved iteratively.

The relationship between collocation at Lobatto points and the Galerkin method is what
motivated Villadsen and Stewart (1967) to select Lobatto points (actually, the equivalent
symmetric problem with polynomials in x2). However, they considered only Dirichlet conditions.
The natural boundary condition treatment for the Galerkin method is advocated by Finlayson
and Scriven (1966). They point out this treatment sets a combination of boundary and interior
residuals to zero, as in Eq. (35). The natural treatment is standard in finite element
applications. The use of this procedure with collocation at Lobatto points was suggested by
Young (1977) and is discussed more fully by Funaro (1992, pp. 143, 204). Canuto (1986)
discussed an apparently similar procedure for Chebyshev and Lobatto points. Canuto et al.
(2006) describe the procedure more clearly, but unfortunately rename it G-NI (Galerkin with
Numerical Integration), which is likely to cause much confusion. Shen, et al. (2011) use the
more appropriate designation of collocation in the weak form. However, all of these texts also
describe the strong, boundary collocation treatment and do not claim a major benefit for a
natural treatment. Consequently, most applications use boundary collocation. We see no
reason to rename the method. The weak formulation should be the standard, correct method,
while the strong treatment is incorrect. We believe the examples below and in section 4
provide sufficient support for use of a natural or weak boundary condition treatment rather than
boundary collocation and justify the words “correct” and “incorrect” used above.

[25]
3.3.1 Linear Source, Third Kind B.C., Galerkin/Moments
Fig. 10 shows the flux errors from Fig. 9 10 0

updated with results from full Galerkin and


-2
10
moments methods and Lobatto and
Chebyshev collocation with a natural 10 -4

treatment of the boundary conditions. These

Flux Error
-6
10
results are labeled “nat.” while the boundary
collocation results from Fig. 9 are labeled 10
-8 Lobatto, nat.
Gauss
“bc”. -10 Cheb. nat.
10 Lobatto, bc
Cheb., bc
The improvement by using natural boundary 10 -12
Galerkin
Moments
conditions is impressive, especially with -14
10
Lobatto points. With the natural boundary 2 4 6 8 10 12 14
n
Fig. 10 Flux Error Eqs. (10) and (10b), Bi = 10  = 5
condition treatment, the convergence rates
for the three choices of points are similar to those in Fig. 6. Since the problem reduces to the
Dirichlet problem for large Bi, the effect of this parameter was investigated by solving the
problem with Bi = 2, 5, 10 and 50 for comparison. With the largest value, the condition
approaches a Dirichlet condition. For the Lobatto cases in Fig. 10 with n = 8, the ratio of the
error with boundary collocation relative to a natural treatment, is 1.4x10 3. This error ratio at n =
8 varies from 7.0x103 when Bi = 2 to 0.3x103 for Bi = 50. The convergence rates are similar, so
the disparity grows with n. Clearly, the error with boundary collocation can be several orders of
magnitude greater even for quite large values of Bi. The errors with Chebyshev points are
reduced to a lesser extent by using a natural boundary condition treatment, but still almost an
order of magnitude at n = 8.

As discussed above, the natural boundary condition treatment, Eq. (35), sets a weighted
average of the boundary and interior residuals to zero. Neither residual will be identically zero,
but they converge to zero at an exponential rate. Fig. 11 shows the convergence behavior of
these two residuals as a function of n for Bi = 10. Graphs for the other values of Bi are similar.
2
10
As shown in Eq. (35), the ratio of the two
residuals is equal to the boundary quadrature 10
0

weight, W0 and Wn+1 or approximately n for


-2
R at boundary
-2
10
Lobatto quadrature and half that for Clenshaw- b.c.residual
Residual

-4
Curtis quadrature (Chebyshev points). 10

-6
10
We also note that in Fig. 10 the Galerkin and Lobatto
-8
10
moments methods improve in a stair step Chebyshev
Gauss
fashion. When n is even, the results with 10
-10

Galerkin and Lobatto points are identical and -12


10
5 10 15 20
the results with moments and Gauss points are n
Fig. 11 Boundary condition residual and interior residual
identical. This result is an artifact caused by the at boundary, Bi = 10,  = 5, k = 1

[26]
symmetry of the solution about x = 0.5. For example, with n = 4 the trial solution is a 5th order
polynomial, but the highest order term has a zero coefficient because of the symmetry. Lobatto
points normally miss exact integration of the Galerkin method by one degree, but since the
problem symmetry knocks out the highest degree, the two methods agree. With n = 3 the
polynomial is 4th order. Adding a 5th order term does not improve the Galerkin solution. With n
= 3 collocation at Lobatto points does not agree with the Galerkin method because the
integration is approximate. An analogous situation applies for the moments method and
collocation at Gauss points.

When the differential equation contains values of the dependent variable (y in this case), not
exclusively its derivatives, this very special set of circumstances, i.e. linear, constant coefficient
and symmetric solution, are essentially the only times when collocation at Lobatto points is
identical to a Galerkin method and collocation at Gauss points is identical to the moments
method. If the source r(x,y) in Eq. (10) is dependent only on x and is a polynomial of degree n
or less, collocation at Gauss and Lobatto points are identical to moments and Galerkin
methods, respectively. Other conditions can easily be evaluated by comparing the accuracy of
the quadrature to the degree of polynomial which must be integrated. Usually, Gauss and
Lobatto points give approximations to the moments and Galerkin methods, but often very good
ones. Some texts incorrectly state that collocation at Gauss points replicates the Galerkin
method for more general conditions [Boyd (2000), p. 89].

3.4 Mass Conservation and Fluxes


One is usually interested in the flux at the boundaries. For example, if a fluid were flowing on
both sides of the slab we would want to know the rate of mass or heat transfer from the slab.
For the symmetric problem, the transfer is quantified by a single normalized flux, η given by
Eqs. (13) and (14). For a nonsymmetric problem one would generally want the breakdown of
left and right side fluxes, while in multiple dimensions the flux profiles along the boundaries
may be important. The divergence theorem, Eq. (14), or average energy equation, gives only
the total flux. If the sum of the individual boundary fluxes (or integral in multiple dimensions)
equals the total flux, this gives us greater confidence in the individual values. Conversely, if the
fluxes do not achieve an overall balance, we have less confidence in the solution.

The total of the fluxes will give the average rate provided the method conserves mass. The
divergence theorem, Eq. (14), is nothing more than an overall balance (of mass, energy, etc.).
It is derived by integrating Eq. (10) across the domain. In general, a method will be
conservative if the integral of the residual is zero. If the MWR weights contain unity in some
combination the method will be conservative.

Moments and collocation at Gauss points are conservative. If formulated with monomials or
Legendre polynomials the first weight function would be x0 or P0 = 1. In our formulation of
section 3.2, ∑ ℓ∗𝑖 (𝑥) = 1. Since these methods are conservative, Eq. (14) is obeyed, so the
same value of the boundary flux is calculated by integration, Eq. (21) or by differentiation of the

[27]
approximate solution, Eq. (22). This explains why the approximations in Fig. 6 are the same
with either method of calculation for collocation at Gauss points.

For the Galerkin method and collocation at Lobatto points, the sum of all the Lagrange
interpolating polynomials is also unity. However, if we consider Dirichlet boundary conditions,
the boundary values are directly substituted into Eq. (2) so the first and last interpolating
polynomials are not used as weight functions. The method appears not to be conservative due
to the leftover terms on the right side below:
𝑑𝑦 1 1 1
| + ∫ 𝑟(𝑥, 𝑦) 𝑑𝑥 = ∫ [ℓ0 (𝑥) + ℓ𝑛+1 (𝑥)]𝑅(𝑥, 𝒚)𝑑𝑥 (36)
𝑑𝑥 0 0 0
Using quadrature, Eq. (36) is approximated by:
𝑑𝑦 1 1
| + ∫ 𝑟(𝑥, 𝑦) 𝑑𝑥 = 𝑊0 𝑅(0, 𝒚) + 𝑊𝑛+1 𝑅(1, 𝒚) (37)
𝑑𝑥 0 0
The two terms on the right side are those needed to correct the fluxes in the natural boundary
condition treatment, Eq. (35), so the correction makes the method conservative. To be
consistent with the Galerkin method, the flux on the left side should be approximated by:
𝑛+1 𝑛+1 𝑛+1
𝑑𝑦
| = ∑ 𝐴0𝑖 𝑦𝑖 + 𝑊0 (∑ 𝐵0𝑖 𝑦𝑖 + 𝑟(0, 𝑦0 )) = ∑ 𝐴0𝑖 𝑦𝑖 + 𝑊0 𝑅(0, 𝒚) (38)
𝑑𝑥 𝑥=0
𝑖=0 𝑖=0 𝑖=0
The same equation written with the stiffness matrix is:
𝑛+1
𝑑𝑦
| = − ∑ 𝐶0𝑖 𝑦𝑖 + 𝑊0 𝑟(0, 𝑦0 ) (39)
𝑑𝑥 𝑥=0
𝑖=0
Analogous equations apply at the right boundary. When computed from these equations, the
two fluxes will be consistent with Eq. (14), the average energy equation or divergence
theorem. For the full Galerkin method, the fluxes must be calculated with the equivalent
expression using the mass matrix, D. For the Robin boundary conditions, the expressions are
consistent with Eq. (10b), so it would be far simpler to calculate fluxes by multiplying 2Bi by the
boundary value, y0 or yn+1. However, that calculation is subject to roundoff errors when Bi is
large.

Note that Eqs. (38) and (39) are also valid with Gauss points, since the boundary quadrature
weights are zero. This equivalence is useful when writing one computer code which will work
with all types of points, provided one is not averse to a few unnecessary calculations. Using
these equations with Chebyshev points will also make that method conservative, which
provides some additional justification for using a natural treatment with Chebyshev points.

The relationship shown above, between the overall balance and natural boundary condition
treatment was noted by Finlayson and Scriven (1966). For flux boundary conditions, Ferguson
(1971) suggested replacing boundary collocation with the integral of the rate, see Eq. (14), to
enforce mass conservation and improve accuracy. This integral procedure has several
limitations. The natural treatment is equivalent, more general and easier to implement.

[28]
4. Parabolic Problem - Falling Liquid Film
In their original article, Villadsen and Stewart (1967) considered two parabolic problems: (1)
the transient version of the catalyst pellet problem and (2) eigenvalues for the classic Graetz
problem. As an example of a parabolic equation we consider a convective mass transfer
problem. The problem is one of mass transfer from a liquid film flowing down a wall. Bird, et al.
(1960, p. 538) describe the problem and present a penetration solution valid near the inlet. The
problem was also treated by Villadsen and Michelsen (1978) using collocation at the base
points of Radau quadrature. We wish to supplement their results by considering other choices
of points, Gauss, Radau, Lobatto and Chebyshev points. The problem has a no flux (Neuman)
condition at one end, so it provides an opportunity to further test different methods for treating
boundary flux conditions.

Consider laminar flow of a liquid film down a solid wall. Let the gas-liquid interface be at x = 0
and the solid wall at x = 1. The film is in laminar flow, so the velocity profile is parabolic, 0 at x
= 1 and a maximum at x = 0. Like most convective heat and mass transfer problems, the no
slip condition at the wall creates a singularity there. Assume the entering liquid is at one
composition and the gas-liquid interface is at a different composition. Since the wall is solid,
assume no flux there. The dimensionless governing equations for the problem are:
2
𝜕𝑦 𝜕 2 𝑦
(1 − 𝑥 ) = (40)
𝜕𝑧 𝜕𝑥 2
with:
𝜕𝑦
𝑦(0, 𝑧) = 0, (1, 𝑧) = 0 and 𝑦(𝑥, 0) = 1 (40a)
𝜕𝑥
The heat transfer analog of the problem is obvious. This model can be compared to a simpler
lumped parameter model where the transfer is governed by a mass transfer coefficient:
2 𝑑𝑦̅
= −𝑆ℎ 𝑦̅ (41)
3 𝑑𝑧
where Sh is the Sherwood number. For heat transfer, the analogous parameter would be the
Nusselt number, a dimensionless heat transfer coefficient. 𝑦̅ is the mixing cup average
composition given by:
1
∫0 (1 − 𝑥 2 ) 𝑦 𝑑𝑥 3 1
𝑦̅(𝑧) = 1   = ∫ (1 − 𝑥 2 ) 𝑦(𝑥, 𝑧) 𝑑𝑥 (42)
∫0 (1 − 𝑥 2 ) 𝑑𝑥 2 0

By using the average energy equation or divergence theorem, the Sherwood number can be
calculated from the solution by two different methods:
−2 𝑑𝑦̅  1  𝜕𝑦
𝑆ℎ = = (0, 𝑧) (43)
3𝑦̅ 𝑑𝑧 𝑦̅ 𝜕𝑥
The Sherwood number is another normalized flux quantity, like the effectiveness factor, η, in
the catalyst pellet problem of Section 3.

[29]
The conventional method of solution by the orthogonal collocation or pseudospectral method is
to use the differentiation matrix to convert the problem to a coupled set of ordinary differential
equations:
𝑛+1
𝑑𝑦𝑗 (𝑧)
(1 − 𝑥𝑗2 ) − ∑ 𝐵𝑗𝑖 𝑦𝑖 (𝑧) = 0 (44)
𝑑𝑧
𝑖=0
for j = 1,…,n, with
𝑛+1

𝑦𝑗 (0) = 1, 𝑦0 (𝑧) = 0, and ∑ 𝐴𝑛+1,𝑖 𝑦𝑖 (𝑧) = 0 (44a)


𝑖=0
As before, A and B give approximations to the first and second derivatives as defined in Eq.
(20).

Eq. (44a) approximates the no flux condition at x = 1 using boundary collocation as suggested
in most texts on the orthogonal collocation or pseudospectral method. Based on the results in
section 3 for the catalyst pellet problem, we can expect the boundary collocation approach to
be inferior to a natural treatment, except for Gauss or Radau-left points since they give a
quadrature weight of zero at the no flux boundary, Wn+1 = 0.

The base points and weights for Gauss, Radau-left, Lobatto and Clenshaw-Curtis quadrature
(Chebyshev points) are compared in Fig. 2. The Radau-left points used by Villadsen and
Michelsen are skewed away from x = 0 and are more densely spaced near x = 1. This choice is
somewhat counterintuitive, since for this problem large gradients occur near x = 0 for small z
and the point spacing is less dense in that area. The Radau-right points are the mirror image of
the Radau-left points and have a greater density of points near x = 0. Although Villadsen and
Michelsen did not state their reasoning, most likely they selected Radau-left points to achieve
the greatest quadrature accuracy with Wn+1 = 0.

As discussed above in section 3, an alternative weak formulation can be constructed using the
stiffness matrix. For Eq. (40) the weak formulation is:
𝑛+1
𝑑𝑦𝑗
(1 − 𝑥𝑗2 )𝑊𝑗 + ∑ 𝐶𝑗𝑖 𝑦𝑖 = 0 (45)
𝑑𝑧
𝑖=1
for j = 1,…,n+1 where C is the stiffness matrix given by Eq. (29). In Eq. (45) we have used the
boundary conditions, Eqs. (40a), which is not obvious since a zero was substituted. By
examining the definition of C it is clear that at the interior points Eq. (45) is equal to Eq. (44)
multiplied by the quadrature weights, Wj, so the equations are equivalent. They differ only at
the boundary point, n+1, where Eq. (45) is equal to:
𝑛+1 𝑛+1

∑ 𝐴𝑛+1,𝑖 𝑦𝑖 − 𝑊𝑛+1 ∑ 𝐵𝑛+1,𝑖 𝑦𝑖 = 0 (46)


𝑖=0 𝑖=0
Eq. (46) differs from Eq. (44a) by the additional term which is again the interior residual
evaluated at the boundary multiplied by the boundary weight, Wn+1. This is the same

[30]
relationship as before, see Eq. (35). With this natural treatment of the boundary condition, we
can expect the residual of the boundary condition to converge along with the residual of the
differential equation as in Fig. 11. So, the no flux boundary condition is embedded in Eq. (45).
The boundary weight is zero for Gauss and Radau-left points, so the two formulations, Eqs.
(44a) and (46), are equivalent for those cases. The stiffness matrix, C, is symmetric for Gauss,
Radau (left and right) and Lobatto points and for Chebyshev points when n < 4.

The weak form, Eq. (45), with Lobatto points approximates the Galerkin method. The stiffness
matrix is exact, but the mass matrix is approximated by the diagonal one on the left side of Eq.
(45). Due to the parabolic velocity profile, Lobatto quadrature misses exact integration of the
Galerkin mass matrix by 3 degrees. A full Galerkin method can be determined by replacing the
diagonal approximation with a more accurate full mass matrix.

For the conventional formulation values, y0 and yn+1, are eliminated using the boundary
conditions so Eq. (44) is reduced to:
𝑛
𝑑𝑦𝑗
(1 − 𝑥𝑗2 ) − ∑ 𝐵̂𝑗𝑖 𝑦𝑖 = 0 (47)
𝑑𝑧
𝑖=1
where:
𝐴𝑛+1,𝑖
𝐵̂𝑗𝑖 = 𝐵𝑗𝑖 − 𝐵𝑗,𝑛+1
𝐴𝑛+1,𝑛+1
Similarly, elimination of boundary values from Eq. (45) gives:
𝑛
𝑑𝑦𝑗
(1 − 𝑥𝑗2 )𝑊𝑗 + ∑ 𝐶̂𝑗𝑖 𝑦𝑖 = 0 (48)
𝑑𝑧
𝑖=1
where:
𝐶𝑛+1,𝑖
𝐶̂𝑗𝑖 = 𝐶𝑗𝑖 − 𝐶𝑗,𝑛+1
𝐶𝑛+1,𝑛+1
for j = 1,…,n

4.1 Continuous Solutions in z


Eqs. (47) or (48) are a coupled set of ordinary differential equations. They would normally be
solved numerically by one of many methods for initial value problems, e.g. Runge-Kutta or
others [Hairer, et al. (1993), Hairer and Wanner (1996)]. Solutions of this type are often called
the Method of Lines, referring to the lines traced out in z by the collocation points in x.
However, to gain greater insight, we will solve the problem analytically. The analytical solution
can be found by assuming a solution of the form 𝑦 = 𝑒 −𝜆𝑧 and then computing the eigenvalues
and eigenvectors of the generalized eigenproblem:
̂ 𝒚 = 𝜆𝑾
𝑪 ̂𝒚 (49)
̂ is replaced by −𝑩
In Eq. (49), 𝑪 ̂ for the first formulation, and 𝑾
̂ is the diagonal matrix
multiplying dy/dz in either formulation.

[31]
If the problem is solved by a full Galerkin method the dense mass matrix prevents the
elimination of the boundary point, yn+1, so the coefficients of dyn+1/dz are nonzero, C is used in
place of 𝑪̂ and the full mass matrix is substituted for 𝑾
̂. Consequently, since i and j = 1,…n +
1 there is an additional eigenvalue with the Galerkin method. Some would call this a spurious
eigenvalue. A dense mass matrix creates two complications for numerical solutions of the
problem: (1) because of the full mass matrix a linear algebra problem must be solved even for
an explicit method and (2) the extra eigenvalue is usually large making the problem stiff so an
explicit method is not appropriate anyway.

The continuous time solution to the problem is [Franklin (1968)]:


𝑛 𝑛 𝑛
−1
𝑦𝑖 = ∑ ∑ 𝑣𝑖𝑘 𝑣𝑘𝑗 𝑦𝑗 (0) 𝑒 −𝜆𝑘𝑧 = ∑ 𝑎𝑖𝑘 𝑒 −𝜆𝑘𝑧 (50)
𝑘=1 𝑗=1 𝑘=1

where vik, are the eigenvectors. Coefficients 𝑎𝑛+1,𝑘 for the boundary value yn+1 are calculated
by substitution of Eq. (50) into Eq. (44a) or (46). Using Eq. (42) the mixing cup composition is:
𝑛 𝑛 𝑛
3
𝑦̅ = ∑ ∑ 𝑊𝑖 (1 − 𝑥𝑖2 ) 𝑎𝑖𝑘 𝑒 −𝜆𝑘𝑧 = ∑ 𝑐𝑘 𝑒 −𝜆𝑘𝑧 (51)
2
𝑘=1 𝑖=1 𝑘=1
Determination of the Sherwood number requires calculation of the boundary flux. As shown in
Eq. (43), the boundary flux can be calculated from either 𝜕𝑦/𝜕𝑥|𝑥=0 or from 𝑑𝑦̅/𝑑𝑧. If
calculated correctly as discussed in the section 3.4, it makes no difference which method is
used. The coefficients of the boundary flux are conveniently given by the weak form of the
approximation:
𝑛+1 𝑛 𝑛 𝑛
𝜕𝑦
| = − ∑ 𝐶0,𝑖 𝑦𝑖 = − ∑ ∑  𝐶0,𝑖 𝑎𝑖𝑘 𝑒 −𝜆𝑘𝑧 = ∑ 𝑑𝑘 𝑒 −𝜆𝑘𝑧 (52)
𝜕𝑥 𝑥=0
𝑖=1 𝑘=1 𝑖=1 𝑘=1
Eq. (52) gives the same boundary flux as the differentiation of Eq. (51). The Sherwood number
is then:
−2 ∑𝑛𝑘=1 𝜆𝑘 𝑐𝑘 𝑒 −𝜆𝑘𝑧 ∑𝑛𝑘=1 𝑑𝑘 𝑒 −𝜆𝑘𝑧
𝑆ℎ = = 𝑛 (53)
3 ∑𝑛𝑘=1 𝑐𝑘 𝑒 −𝜆𝑘𝑧 ∑𝑘=1 𝑐𝑘 𝑒 −𝜆𝑘𝑧
The asymptotic Sherwood number applies for large z:
−2𝜆1
𝑆ℎ∞ = lim 𝑆ℎ = (54)
𝑧→∞ 3

The analytical solution to the problem is like Eq. (51) except that it is an infinite series. The
collocation solution approximates a truncation of the infinite series. In the analytical solution, all
the eigenvalues are not only real and negative but also the coefficients, c, are all positive. The
solution is a series of different modes each decaying at a different rate.

The statements above apply to the analytical solution, but what about the numerical solutions?
Whether an approximation produces real eigenvalues and positive values for c, is one test for

[32]
the suitability of an approximate method. When 𝑪 ̂ and 𝑾 ̂ in Eq. (49) are symmetric and
positive definite the eigenvalues are all real and positive. The weak formulation, Eqs. (45) and
(48), produces symmetric positive definite matrices for all but Chebyshev points with n > 3. As
explained above, the conventional formulation, Eqs. (44) and (47), and weak formulation are
equivalent when the boundary quadrature weight is zero, Wn+1 = 0, which is true for Gauss and
Radau left points. Boundary collocation makes the matrix asymmetric for Lobatto and Radau
right points and creates additional asymmetry for Chebyshev points. The consequences of this
asymmetry will be discussed shortly.

An eigenvalue problem like this can be solved more efficiently when matrices are symmetric.
When the matrices are not symmetric there is no advantage to treating the eigenproblem in
̂, to give the
generalized form, Eq. (49), so one should divide through by diagonal matrix, 𝑾
identity matrix on the right hand side.

To test for possible complex eigenvalues and negative coefficients, c, all the nonsymmetric
eigenproblems were solved for n ≤ 60. For the weak formulation, Chebyshev points produce
only real eigenvalues and positive coefficients, c. For the conventional formulation, Eqs. (44)
and (47), Radau-right points give complex eigenvalues for n > 4, Lobatto points give complex
values for n = 4-6 and n > 11, Chebyshev points give complex values for n = 28, 45, 50 and
58. The coefficients, c, are negative for many cases when the eigenvalues are real. The
propensity to produce complex eigenvalues and negative coefficients is directly related to the
magnitude of the boundary quadrature weights, Wn+1, and hence the degree of difference
between Eqs. (44a) and (46). For large n, the Chebyshev (Clenshaw-Curtis) boundary weights
are half the Lobatto and Radau boundary weights.

On the surface, these results appear to be different from previous work which has analyzed
approximations to the constant coefficient heat equation. Gottlieb and Lustman (1983) found
that Chebyshev points produce only real eigenvalues for the heat equation with general
boundary conditions that include those in Eq. (40a). The analysis of the heat equation by
Canuto, et al. (1988, p.407) proved real eigenvalues for Lobatto and Chebyshev points with
Neumann and Dirichlet boundary conditions. However, they consider a weak formulation with a
natural boundary condition treatment, like that used here. Funaro (1992, pp. 143, 204) also
considered eigenvalues with a natural boundary condition treatment. Unfortunately, these
previous works are frequently cited to support the suitability of the conventional formulation
using boundary collocation. It is obviously incorrect to generalize the results to a different
boundary condition treatment, and in this case to a problem with variable coefficients

To reconcile these differences, we did some limited testing of the constant coefficient heat
equation with the more common boundary collocation treatment. We found that with boundary
collocation, Chebyshev points produced only real negative eigenvalues, but the coefficients, c,
are frequently negative. For Lobatto points, the eigenvalues were complex for n > 6, so with

[33]
Lobatto points the conventional formulation with boundary collocation is unsuitable even for the
constant coefficient problem.

The accuracy of our results were double checked with two completely independent
calculations: (1) in Fortran using the LAPack code DGEEV [Anderson, et al. (1999)] and (2) in
Octave using the eig function and code built upon that of Trefethen (2000). The results agreed
to within the limits of machine precision (about 13 digits).

Tables 2 and 3 summarize this discussion concerning the suitability of the various
formulations. Based on these results, we conclude that Lobatto and Radau-right points are
unsuitable using the conventional formulation with boundary collocation but are suitable using
the weak formulation. We conclude the conventional formulation with Chebyshev points
(excluding n = 28, 45, 50 and 58) is marginally acceptable. The other formulations are
suitable, especially those that produce symmetric matrix problems.
Table 2 Table 3
Conventional Formulation, Eq. (44) Weak Formulation, Eq. (45)
Equivalent Points Symmetric Eigenvalues
Points Eigenvalues
To Eq. (45) Gauss yes real
Gauss yes real Chebyshev no real
Chebyshev no complex Lobatto yes real
Lobatto no complex Radau Left yes real
Radau Left yes real Radau Right yes real
Radau Right no complex

Fig. 12 shows the first few eigenvalues for n = 6. The notation “bc” (boundary collocation)
designates the conventional formulation of the Chebyshev method, i.e. Eq. (44) and (44a). Fig.
13 shows the convergence of the first and fourth eigenvalues. The convergence of the
coefficients mirrors that of the eigenvalues, so are not presented. The relative convergence
rates of the various points are like those for the catalyst pellet problem. However, in this case
we have added the Radau points, which tend to have accuracy between that of Lobatto and
Gauss points. Chebyshev points with boundary collocation give the worst convergence, which
is disconcerting since it is probably the most popular formulation. Although no formal analysis
was performed, these calculations appear to reach the limits of roundoff with about 13 digits of
accuracy. Figs. 12 and 13, show that the first n/2 eigenvalues are usually accurate and the
higher eigenvalues are greater than the actual ones.

Villadsen and Michelsen give asymptotic relationships for the eigenvalues and coefficients of
the analytical solution. From information given there and the approximate solutions the
following asymptotic relationships were determined:
𝜆𝑘 → (4𝑘 − 1.672)2
(55)
𝑐𝑘 → 3.820/𝜆𝑘

[34]
0
10
Lobatto
104 Exact
-2
Gauss
Chebyshev bc 10 Radau L
Chebyshev Radau R
Radau Right Chebyshev
-4 Chebyshev bc
Radau Left 10 Galerkin
3
10
eigenvalue

Lobatto
Gauss

error
Galerkin 10-6
4
2
10 10 -8

1
-10
10
1
10
10-12
1 2 3 4 5 6 7 5 10 15 20 25
n n
Fig. 12 Eigenvalues calculated with n = 6 Fig. 13 Error in first and fourth eigenvalues
These relationships give values within 0.1% for k = 6 to 51 of the values calculated with 100
Lobatto points.

Fig. 12 shows that the higher eigenvalues 10


6

Lobatto
are greater than the actual ones. The Gauss
105
magnitude of the largest eigenvalues is of Chebyshev
Radau Left
interest when explicit numerical integration 4
Radau Right
Max Eigenvalue

10 Galerkin
methods are considered. For these (4n-1.672)
2

methods, the maximum step size is 103


inversely proportional to the magnitude of
the maximum eigenvalue. Previous work 10
2

has found that asymptotically, the


maximum eigenvalues for the constant 101

coefficient heat equation increase with n4


0
10
[Canuto, et al. (2006)]. Fig. 14 shows the n 5 10 15
Fig. 14 Maximum eigenvalues vs. number of points
first 15 maximum eigenvalues for this
problem. For those shown, the maximum values are up to two orders of magnitude larger than
the actual ones. Larger eigenvalues have been calculated and those for n = 30 - 100 fit to the
following equation:
𝑛 𝑏
𝜆𝑚𝑎𝑥 → 𝑎 ( ) (56)
30
Values for Eq. (56) are given in Table 4. It appears that if still larger values were fit, the true
asymptote would have an exponent of 6. The additional power of 2 relative to the constant
coefficient problem is to be expected from the norm of 𝑾 ̂, given the point spacing is O(n-2)
near the boundaries. As a result, the eigenvalues for large n are much larger than those for the
constant coefficient heat equation. The eigenvalues of Radau points were not fit, since for
large n the values for Radau left points follow those for Gauss points and the values for Radau
right points follow those for the Lobatto points.

[35]
Table 4 Parameters of Eq. (56) For this problem, explicit numerical
Lobatto Gauss Chebyshev Galerkin integration methods in z are only practical for
a 4.05x10 18.3x10
6 6
7.98x106 62.5x106 relatively small n. If for larger n, the step size
b 5.822 5.932 5.869 5.690 is limited by stability considerations, then
relative to Lobatto points, Chebyshev and
1 20
Gauss points require 2 and 4.5 times as 18
0.8 Lobatto 16
many steps, respectively. 14
Gauss
0.6 Chebyshev 12

Although Eq. (56) is interesting, it is of little Radau Left


10
Radau Right
0.4 8
practical consequence if only a few points are exact

yavg

Sh
required to give the desired accuracy. The 6

real power of these methods is that they


frequently produce good accuracy with only a 0.2 4

few points. For example, Fig. 15 shows y̅


and Sh as a function of axial position for the
various point choices with n = 2. All the 0 0.1 0.2 0.3
2
0.4
z
higher terms damp out by z = 0.10, so the Fig. 15 Average composition and Sh vs z, n = 2
solution is asymptotic from there on. The
slope and intercept of the asymptotic part is determined by the first eigenvalue and coefficient,
respectively. Since one would normally be interested in transferring all or most of a
component, the figure shows the portion of the curve for approximately 90 percent transfer
which is reached at z = 0.4. Except for Gauss points, all the methods give good results for z >
0.05.
1.2
If one is interested in accurate solutions for
very small z then larger n is required. The 1

behavior at small z is much like that of an


0.8
infinite series analytical solution when the
series is truncated as it must be for practical Lobatto n = 3
0.6
y

Gauss n = 3
calculations. Fig. 16 shows the profiles at z = Chebyshev n = 3
0.01 for n = 3. For the solutions in Fig. 16, 0.4 Radau Lft n = 3
Radau Rgt n = 3
1% is a typical error in average composition. exact
0.2
Contrary to intuition, the Radau-right points
are not appreciably more accurate even
0
0 0.2 0.4 0.6 0.8 1
though the points are more densely spaced x
Fig. 16 Calculated profiles at z = 0.01, n = 3
near x = 0.

Larger n is required to determine accurate flux or Sh values at small z. Fig. 17 shows the
calculation of Sh for Lobatto points with different n. From a penetration solution valid for small
z, the Sherwood number is proportional to 1/√𝑧. Any polynomial solution will eventually
breakdown as one moves toward the discontinuous initial profile, so all the different points

[36]
display the general behavior seen in Fig. 60
50
17. However, there are differences. At 40
small z, some tend to level out at larger Sh 30
Lobatto n = 2
Lobatto n = 4
and have a larger maximum positive Lobatto n = 6
20
deviation after crossing over at larger z. For Lobatto n = 8
Lobatto n =10
a range of n, the maximum positive

Sh
exact
deviations are 4 to 5 percent for Lobatto 10

and Radau-left points, 25 to 35 percent for


Gauss and Radau-right points and 10 to 15
percent for Chebyshev points. This
behavior can be seen in Fig. 15 to some
10-4 10-3 10-2 10-1 100
extent. z
Fig. 17 Sh vs z, Lobatto points
0
Fig. 18 shows the overall error in the 10

calculated flux versus n. The errors in y̅ Lobatto


10
-2 Gauss
looks similar, but at a given n are 5-10 Radau L
Radau R
times smaller at z = 0.10 and 10-50 times 10
-4
Chebyshev
smaller at z = 0.01. Fig. 18 looks like Fig. Chebyshev bc
error flux

-6
13, but the curves are more tightly clustered 10
z = 0.01
for the various points. For z > 0.01 and n = -8
10
6, the error is typically less than 1% in the
flux and 0.05% in y̅ . The errors in y̅ are 10
-10

typically less than 0.1% for n = 3 and z > z = 0.10


0.1. These results indicate that the behavior 10
-12
5 10 15 20 25 30
n
of the eigenvalues (Eq. (56), Table 3) at Fig. 18 Flux error at z = 0.01 and 0.10
large n is of no practical importance unless
one is only interested in the behavior at very small z. For these conditions, a finite element
method is better suited.

4.2 Summary Remarks - Global Methods and Fluxes


The conventional formulation using a strong, boundary collocation treatment for flux conditions,
Eqs. (19) and (44a), produces poor approximations except for cases when the quadrature
weight is zero at the flux boundary (see Appendix A). With Lobatto and Chebyshev points, the
results are bad enough for the boundary value problem in section 3, the parabolic problem in
section 4, and other problems [Ferguson (1971), Finlayson (1972), Elnashaie and Cresswell
(1973), Young (2018)], that they can be called wrong. Their use for parabolic problems often
produces nonphysical complex eigenvalues. Even when the eigenvalues are not complex, it
gives the slowest rate of convergence. Errors are typically several orders of magnitude greater,
see Figs.10,13 and 18. Some problems show significant errors in both the internal profiles and
fluxes, while for others only the fluxes show large errors. The weak formulation is the correct
treatment for Lobatto and Chebyshev points, e.g. Eqs. (33), (35), (45) or (46). In addition to

[37]
greater accuracy, the weak formulation with natural boundary conditions is more efficient due
to symmetric matrices for all but Chebyshev points.

For flux boundary conditions, boundary collocation is the correct approach with Gauss points
(or Radau points with a weight of zero at the flux boundary). If one prefers the more intuitive
boundary collocation treatment, then one should follow the long-standing recommendation and
use Gauss points [Young and Finlayson (1976), Michelsen and Villadsen (1981)]. Collocation
at Gauss points works quite well and is the most popular choice in OC applications. This
conclusion is contrary to the oft stated claim that nonzero boundary weights somehow facilitate
the approximation of boundary conditions.

If the boundary conditions are correctly treated, collocation at Lobatto points is usually slightly
more accurate than collocation at Gauss points, i.e. equivalent to about one increment of n.
Collocation at Lobatto points is somewhat better suited for explicit time stepping methods due
to smaller more accurate eigenvalues, but if large n is required, explicit methods are
unsuitable in any case. Collocation at Chebyshev points usually produces somewhat less
accurate flux calculations and can be less efficient because matrix problems are always
nonsymmetric. Collocation at Chebyshev points appears not to be competitive with the other
choices unless the problem requires large n and is amenable to FFT calculations.

5. Multidimensional Applications
The global orthogonal collocation methods described in the forgoing examples are readily
extended to two dimensions using tensor product trial functions, Eq. (3). An example for the
Poisson problem can be found in Villadsen and Stewart (1967). This author’s first application
of the method was in 1971 for a class assignment to solve the problem in section 3 for finite
cylindrical catalyst pellets, i.e. r-z coordinates. This problem was later the subject of a technical
paper [Sorensen, et al. (1973)]. Numerous other examples are available in the literature [e.g.
Michelsen and Villadsen (1972,1981), Finlayson (1974, 2003), texts listed in section 2.4].

Global methods have also been applied to distorted domains, which can be mapped to
rectangles. This approach was used by Young and Finlayson (1976) and Orzag (1980), and is
fully discussed in Canuto, et al. (2006). Although these methods permit some geometric
flexibility, finite element based methods have much greater flexibility.

6. Orthogonal Collocation - Finite Elements


The examples above used global polynomial trial functions, i.e. a single polynomial represents
the solution. Although the global orthogonal collocation method works quite well for many
problems, it has difficulties when the solution cannot be accurately represented by a single low
order polynomial, e.g. the catalyst pellet problem of section 3 with large φ or the falling film
problem in section 4 for small z. In addition, finite element methods have greater flexibility for
representing irregularly shaped domains. There was an explosion of interest in finite elements
in the 1970’s [Zienkiewicz (1971), Strang and Fix (1973), Hughes (1987)]. Their success in
[38]
structural mechanics led to interest in other disciplines. It is not surprising that soon after the
Villadsen and Stewart article, collocation methods were extended to finite element trial
functions using the same idea of collocating at quadrature points. There were several related
developments in this area that were made independently.

A finite element method differs from a global method by partitioning the domain and using a
different polynomial in each subdomain or element. If a single element is used, the two
approaches are the same, so the finite element approach is more general. With a finite
element approximation, one also has some flexibility in the degree of continuity at the
boundary or interfaces between elements. The notation Cn is used to describe the continuity,
where n denotes the highest derivative which is continuous.

To develop a finite element approximation in one dimension, we first lay out a grid:
𝑥0 < 𝑥1 < ⋯ < 𝑥𝑘 < ⋯ < 𝑥𝑛𝑒 (57)
and define element size as Δ𝑥𝑘 = 𝑥𝑘 − 𝑥𝑘−1. Within each element a local coordinate is used,
𝜉 = (𝑥 − 𝑥𝑘−1 )/Δ𝑥𝑘 . The nodes within each element are at the polynomial roots or quadrature
points, ξi, so a specific node is given by 𝑥𝑖𝑘 = 𝑥𝑘−1 + Δ𝑥𝑘 𝜉𝑖 . The double and single subscript
notation are related by 𝑥𝑘 = 𝑥𝑛+1,𝑘 = 𝑥0,𝑘+1 .

We are interested in methods which are directly related to the global collocation methods
discussed above and will subdivide these methods into those which are C1, i.e. continuous first
derivatives, and those that are C0, i.e. simple continuity. With finite elements, the interface
between elements is an internal boundary. The approximate solution and the flux should be
continuous at these internal boundaries. The discussion of flux boundary condition treatment
above applies directly to the approximations used for these internal boundary conditions. For a
Galerkin method, the interelement flux conditions are usually treated in weak form like natural
boundary conditions.

If the conditions at the element interfaces are strongly or exactly enforced, then:
𝑦𝑛+1,𝑘 = 𝑦0,𝑘+1 and
𝑛+1 𝑛+1
1 1 (58)
∑ 𝐴𝑛+1,𝑖  𝑦𝑖𝑘 = ∑ 𝐴0𝑖  𝑦𝑖,𝑘+1
Δ𝑥𝑘 Δ𝑥𝑘+1
𝑖=0 𝑖=0
If there is a material change at the interface with an abrupt change in say thermal conductivity,
then Eq. (58) would be modified to reflect continuity of the flux. This strong enforcement of
interface conditions is equivalent to boundary collocation. An alternative to the flux condition in
Eq. (58) is to use trial functions with built in continuity. For cubic trial functions, Hermite cubic
polynomials could be used [Hildebrand (1987), p. 282]. For higher order polynomials, a grid
like that in Fig. 19 could be used, where the function and its derivative are specified at the
element boundaries (double nodes) and the function values are interpolated at the single
nodes. With built in continuity, the number of equations and unknowns is reduced by ne – 1
relative to explicit enforcement with the side conditions, Eq. (58). There are other tradeoffs

[39]
between the two approaches. These X Gauss Pts.
Single Nodes
elements easily extend to higher order Elements, double node
elements and multiple dimensions.
X X X X X X X X X X X X
1
Fig. 19 Four C Quartic Elements
6.1 Quadrature in Finite Element Methods
From the forgoing discussion it is clear orthogonal collocation is successful because it is an
accurate approximation to integrated MWR methods, while simpler to apply. Following
Villadsen and Stewart’s (1967) idea, the key is to place the collocation points at quadrature
base points to achieve an accurate approximation to the integrals in either the moments
method using Gauss points or the Galerkin method using Lobatto points. Recall from sections
3.2 and 3.3, that with orthogonal collocation, the integration is one degree shy of that required
for a full moments or Galerkin method. For finite elements, is the same level of accuracy
sufficient? To answer this question, we review the integration accuracy requirements for finite
element methods.

For a general purpose code, analytical integrals are cumbersome to implement, especially for
distorted isoparametric elements, so most codes use numerical integration. Early applications
used Gaussian quadrature, since it gives the highest accuracy for a given number of function
evaluations. These quadrature calculations are time consuming, especially for nonlinear time-
dependent problems which required many such calculations. Collocation seemed like an
obvious way to circumvent these problems, since integration is not required. Many studies
have compared the computational cost of conventional and collocation-like methods [Young
(1977), Gray and Van Genuchten (1978), Houstis, et al. (1978), Finlayson (1979), Frind and
Pinder (1979), Dyksen, et al. (1984), Melenk, et al. (2001), Schillinger, et al. (2015)].

The computational cost depends on the trial functions selected, the quadrature method and the
number of quadrature points used. An obvious question is – how accurate does the integration
have to be? To describe the integration issues, consider a general mass matrix, like Eq. (28):
1
𝐷𝑘𝑗 = ∫ 𝑤𝑘 (𝑥)𝜓𝑗 (𝑥)𝑑𝑥 (59)
0
where the wk are the weight functions and the ψj are the trial functions. We say the method
employs full integration if the quadrature is accurate enough to integrate D exactly. With this
definition, the integration may still be approximate since the equation could have nonlinear
terms or variable coefficients. We call it reduced integration if the integration is not accurate
enough to integrate D exactly.

Several studies have investigated the question of integration requirements for Galerkin finite
element methods [Ciarlet and Raviart (1972), Fix (1972), Strang and Fix (1973), Raviart
(1973), Fried (1974), Ciarlet (1978)]. Although there are some side issues related to problem
smoothness, these studies show that for most problems the maximum rate of convergence can
be achieved with less than full integration. Exact integration of all but the highest degree terms
in Eq. (59) is sufficient to maintain the maximum convergence rate. For example, cubic

[40]
elements can produce a 4th order convergence rate, i.e. the same as for interpolation of a
function. For a Galerkin method, full integration would require an exact integration of all terms
through 6th degree, since both terms in Eq. (59) are cubic. However, the 4th order convergence
rate is maintained when only terms through 5th degree are integrated exactly. 5th degree
integration accuracy is achieved with 3 Gauss points or 4 Lobatto points.

Although the theoretical analyses cited above apply only for Galerkin methods, it is obvious
that the same rule applies to the moments method. A moments method with cubic trial
functions would use discontinuous linear weight functions, so full integration would require
exact integration through 4th degree. To achieve a 4th order convergence rate only terms
through 3rd degree must be integrated exactly, requiring 2 Gauss points or 3 Lobatto points.
This illustration is for cubics, but for higher order trial functions the quadrature keeps pace with
the required integration accuracy. If n is incremented by one to increase the degree of the trial
function, the integrand in Eq. (59) and the quadrature accuracy both increase by two, so the
integration is one degree less than full integration for any n.

For time dependent finite element problems, a related issue is the form of the mass matrix.
When the problem is amenable to explicit time stepping methods, a diagonal or lumped mass
matrix is preferred. If the mass matrix is not diagonal, the algebraic problem for each time step
is like that for an implicit method. Recall from sections 3.2 and 3.3 that global orthogonal
collocation produces a lumped mass matrix due to the node placement and quadrature. Fried
and Malkus (1975) found the same result could be achieved for C0 Lagrange trial functions by
using Lobatto quadrature and placing the nodal interpolation points at the quadrature points.
Furthermore, this procedure produced no loss of convergence rate. Here again, the integration
accuracy is one degree less than that needed for full integration. Apparently, Fried and Malkus
did not consider integrating the stiffness matrix with the same reduced quadrature scheme as
was done in later work.

6.2 C1 Orthogonal Collocation at Gauss Points


Section 3.2 shows that collocation at Gauss points approximates the moments method. The
moments method extends naturally to finite elements with residual weighting by discontinuous
C-1 piecewise polynomials. The interface conditions must be strongly enforced as in Eq. (58) or
with a basis having built in C1 continuity, as in Fig. 19. If the integrals in the moments method
are approximated by n point Gaussian quadrature, full integration is missed by one degree.
From the discussion in section 6.1, this integration is accurate enough to maintain the
maximum possible convergence rate, and it reduces to collocation at Gauss points. Collocation
at Gauss points was described by DeBoor and Swartz (1973) and Douglas and Dupont (1973).
Carey and Finlayson (1975) applied the method using Lagrange trial functions with the
conditions of Eq. (58) satisfied explicitly. Several texts describe the method [Davis (1984),
Lapidus and Pinder (1999), Finlayson (2003)].

[41]
We also note the method of moments is related to the H1 Galerkin method [Douglas, et al.
(1974)]. The H1 Galerkin method is not a Galerkin method in the classic sense, because the
trial and weight functions are different. The H1 Galerkin method weights the residual by the
second derivative of the trial functions, so it is a classic moments method as described by N.M.
Krylov (1926), see section 2.2. When the trial functions are C1, the weight functions are C-1,
discontinuous piecewise polynomials. However, the weight functions for the moments method
span only one element, which contributes to improved efficiency. Since moments and the H1
Galerkin methods are equivalent, C1 Orthogonal Collocation at Gauss points approximates the
H1 Galerkin method.

6.3 C0 Orthogonal Collocation at Lobatto Points


Collocation at Lobatto points has also been extended to finite elements. First, there is the
Hybrid-Collocation-Galerkin method [Diaz (1975), Dunn and Wheeler (1976), Wheeler (1977)].
It is a C0 procedure using the Lagrange trial functions shown in Fig. 20. Collocation is applied
at the internal Lobatto points. At the interface nodes, a Galerkin-like integral condition is used
with weighting by the linear hat functions, top of Fig. 20. This approach is reminiscent of
Ferguson’s (1971) integral approach (see section 3.4) for treating flux boundary conditions in
global collocation.

Another approach, developed independently and usually considered to be fundamentally


different, is described by Gray (1977), Young (1977, 1981), Hennart (1982). The method
developed later by Leyk (1986, 1997) appears to be the same. This method uses a plain
vanilla Galerkin method with Lagrange trial functions like Fig. 20 but with integration performed
using Lobatto quadrature. The interpolation points are coincident with the quadrature points as
in global orthogonal collocation and the finite element procedure of Fried and Malkus (1975).
However, Fried and Malkus used Lobatto quadrature only for the mass matrix, while here all
terms are integrated with the same reduced quadrature procedure. The quadrature misses full
integration by one degree, so based on the discussion
in section 6.1, it is accurate enough to maintain the
maximum possible convergence rate. This procedure
was called the Lobatto-Galerkin method by Young.
Gray and Hennart considered only the quadratic case
with integration by Simpson’s rule, while Young
considered trial functions of any order. This method
was developed again later and renamed the hp
Spectral Element method by Maday and Patera
(1989). The earlier work is not referenced in the
spectral literature; Maday and Patera are credited with
discovery of this popular method [Vosse and Minev
(1996), Canuto, et al. (2007), Karniadakis and
Sherwood (2013)].
Fig. 20 Finite element Lagrange trial
functions [from Young (1981)]

[42]
This method is a straight forward extension to finite elements of the ideas in Section 3.3. There
we showed that at the interior points, the weak form, Eq. (33), is equivalent to collocation, Eq.
(18). The same relationship holds at the interior points with finite elements. We will also show
that at the element interfaces, the hybrid-collocation-Galerkin method and Lobatto-Galerkin or
hp spectral element methods are equivalent. However, the Lobatto-Galerkin or hp spectral
element method gives a matrix structure which is better suited for calculations.

Once again consider the catalyst pellet problem. Convert Eq. (33) to an element k:
𝑛+1
𝑑𝑦 𝑑𝑦 1
𝛿𝑗,0 | − 𝛿𝑗,𝑛+1 | + ∑ 𝐶𝑗𝑖 𝑦𝑖𝑘 − Δ𝑥𝑘 𝑊𝑗  𝑟(𝑥𝑗𝑘 , 𝑦𝑗𝑘 ) = 0 (60)
𝑑𝑥 𝑥0,𝑘 𝑑𝑥 𝑥𝑛+1,𝑘 Δ𝑥𝑘
𝑖=0
At the interior nodes, the first two terms of Eq. (60) are zero and by substitution of Eq. (34) it is:
𝑛+1
𝑊𝑗
∑ 𝐵𝑗𝑖 𝑦𝑖𝑘 + Δ𝑥𝑘 𝑊𝑗  𝑟(𝑥𝑗𝑘 , 𝑦𝑗𝑘 ) = Δ𝑥𝑘 𝑊𝑗 𝑅𝑗𝑘 = 0 (61)
Δ𝑥𝑘
𝑖=0
for j = 1,…,n. Rjk designates the residual at node j in element k, so Eq. (60) is clearly
equivalent to collocation at the interior nodes, just like in global methods, see Eq. (18).

At the element interface nodes, xk, the weight function spans both elements (see Fig. 20), so
the equation for the last node of element k and the first node of k+1 must be combined:
𝑛+1
𝑑𝑦 1
− | + ∑ 𝐶𝑛+1,𝑖 𝑦𝑖𝑘 − Δ𝑥𝑘 𝑊𝑛+1  𝑟(𝑥𝑛+1,𝑘 , 𝑦𝑛+1,𝑘 )
𝑑𝑥 𝑥𝑛+1,𝑘 Δ𝑥𝑘
𝑖=0
𝑛+1 (62)
𝑑𝑦 1
+ | + ∑ 𝐶0,𝑖 𝑦𝑖,𝑘+1 − Δ𝑥𝑘+1 𝑊0  𝑟(𝑥0,𝑘+1 , 𝑦0,,𝑘+1 ) = 0
𝑑𝑥 𝑥𝑜,𝑘+1 Δ𝑥𝑘+1
𝑖=0
Using the standard natural boundary condition treatment at the interface, the flux terms cancel
and we are left with:
𝑛+1 𝑛+1
1 1
∑ 𝐶𝑛+1,𝑖 𝑦𝑖𝑘 + ∑ 𝐶0,𝑖 𝑦𝑖,𝑘+1 − (Δ𝑥𝑘 𝑊𝑛+1 + Δ𝑥𝑘+1 𝑊0 ) 𝑟(𝑥𝑛+1,𝑘 , 𝑦𝑛+1,𝑘 ) = 0 (63)
Δ𝑥𝑘 Δ𝑥𝑘+1
𝑖=0 𝑖=0
The conditions at the interface, Eq. (58), are strongly imposed for the values, but not the flux.
As with boundary flux conditions in the global methods, Eq. (63) is equivalent to setting a
weighted average of residuals to zero. This relationship is made evident by the substitution of
Eq. (34) into Eq. (63):
𝑛+1 𝑛+1
1 1
( ∑ 𝐴𝑛+1,𝑖 𝑦𝑖𝑘 − ∑ 𝐴0,𝑖 𝑦𝑖,𝑘+1 ) − Δ𝑥𝑘 𝑊𝑛+1 𝑅𝑛+1,𝑘 − Δ𝑥𝑘+1 𝑊0 𝑅0,𝑘+1 = 0 (64)
Δ𝑥𝑘 Δ𝑥𝑘+1
𝑖=0 𝑖=0
The term within the large parenthesis is the residual of the interface flux condition, Eq. (58),
while the terms on the right are the interior residuals at the interface, i.e. R as defined in Eq.
(61). The value of the residual on both sides of the interface appear, since the residual is
discontinuous. This result is analogous to Eqs. (35) and (46). The method is still basically
collocation, but in this case a weighted average of residuals is set to zero. The residual of the

[43]
interface condition and both interior residuals will be driven to zero as the grid (either n or ne)
is refined.

Since Eq. (60) is equivalent to collocation at the interior nodes, the only difference between
this procedure and the hybrid-collocation-Galerkin procedure is the treatment of the element
interface. The hybrid method uses the linear hat functions to span the interface while the
Galerkin method uses the endpoint Lagrange polynomials shown in Fig. 20, resulting in Eq.
(63). We will show that since the linear basis is included within the Lagrange basis, the hybrid
method is equivalent.

For a linear source function, the residual is a polynomial of degree n + 1, so the linear
weighting of the hybrid method gives integrals of degree n + 2. Lobatto quadrature, exact for
2n + 1, is more than sufficient, so it is used for all integrals. We make use of the following
relationships which come from the treatment of the second order term in the hybrid method.
Integration by parts gives:
1
𝑑 2 ℓ𝑖 𝑑ℓ𝑖 1 1
𝑑ℓ𝑖 𝑑ℓ𝑖
∫ 𝜉 2
𝑑𝜉 = 𝜉 | − ∫ 𝑑𝜉 = | − 𝛿𝑛+1,𝑖 + 𝛿0,𝑖 or
0 𝑑𝜉 𝑑𝜉 0 0 𝑑𝜉 𝑑𝜉 𝑛+1
𝑛+1 𝑛+1 (65)
∑ 𝜉𝑗 𝑊𝑗 𝐵𝑗𝑖 = 𝐴𝑛+1,𝑖 − ∑ 𝑊𝑗 𝐴𝑗𝑖 = 𝐴𝑛+1,𝑖 − 𝛿𝑛+1,𝑖 + 𝛿0,𝑖
𝑗=0 𝑗=0
Eq. (65) is for the left half of the hat function. An analogous equation applies to the right half.

For the hybrid method, the interface equation in weak form, i.e. after integration by parts, is:
𝑛+1 𝑛+1
1 1 1 1
1 𝑑ℓ𝑖 1 𝑑ℓ𝑖
∑ 𝑦𝑖𝑘 ∫ 𝑑𝜉 − ∑ 𝑦𝑖,𝑘+1 ∫ 𝑑𝜉 − Δ𝑥𝑘 ∫ 𝜉𝑟𝑘 𝑑𝜉 − Δ𝑥𝑘+1 ∫ (1 − 𝜉)𝑟𝑘+1 𝑑𝜉 = 0 (66)
Δ𝑥𝑘 0 𝑑𝜉 Δ𝑥𝑘+1 0 𝑑𝜉 0 0
𝑖=0 𝑖=0

where rk designates the source function in element k. The integrals can be simplified with Eq.
(65); however, to show equivalence, we undo the integration by parts reflected in Eq. (66).

After substitution from Eq. (65) for the left element and its analog for the right one, Eq. (66)
becomes:

𝑛+1 𝑛+1 𝑛+1 𝑛+1


1 1
∑ 𝑦𝑖𝑘 (𝐴𝑛+1,𝑖 − ∑ 𝜉𝑗 𝑊𝑗 𝐵𝑗𝑖 ) − ∑ 𝑦𝑖,𝑘+1 (𝐴0,𝑖 + ∑(1 − 𝜉𝑗 )𝑊𝑗 𝐵𝑗𝑖 )
Δ𝑥𝑘 Δ𝑥𝑘+1
𝑖=0 𝑗=0 𝑖=0 𝑗=0
𝑛+1 𝑛+1
(67)
−Δ𝑥𝑘 ∑ 𝜉𝑗 𝑊𝑗 𝑟𝑗𝑘 − Δ𝑥𝑘+1 ∑(1 − 𝜉𝑗 )𝑊𝑗 𝑟𝑗,𝑘+1 = 0
𝑗=0 𝑗=0

After collecting the residual terms in Eq. (67) we have:


𝑛+1 𝑛+1 𝑛+1 𝑛+1
1 1
(
Δ𝑥𝑘
∑ 𝐴𝑛+1,𝑖 𝑦𝑖𝑘 −
Δ𝑥𝑘+1
∑ 𝐴0,𝑖 𝑦𝑖,𝑘+1 ) − Δ𝑥𝑘 ∑ 𝜉𝑗 𝑊𝑗 𝑅𝑗𝑘 − Δ𝑥𝑘+1 ∑(1 − 𝜉𝑗 )𝑊𝑗 𝑅𝑗,𝑘+1 = 0 (68)
𝑖=0 𝑖=0 𝑗=0 𝑗=0

[44]
Since the interior residuals are zero by Eq. (61), we are left with only interface values, so Eq.
(68) is identical to Eq. (64) above. The hybrid-collocation-Galerkin method [Diaz (1975), Dunn
and Wheeler (1976), Wheeler (1977)] is equivalent to the methods of Gray (1977), Young
(1977), Hennart (1982) and Leyk (1986, 1997), Maday and Patera (1989), Vosse and Minev
(1996), Karniadakis and Sherwood (2013). Previously, these methods were considered
different. Given this equivalence the convergence and super-convergence properties
(discussed below) for the hybrid method apply equally to all. Although this demonstration of
equivalence is for a specific problem, it generalizes to other equations, because the linear hat
functions are included within the Lagrange basis.

Since this one basic procedure has gone by several different names, we ask what is a logical
name for it? Is it orthogonal collocation or a Galerkin method with specific quadrature and
nodal locations? The relationship between the Galerkin method and orthogonal collocation at
Lobatto points has been known since Villadsen and Stewart (1967) and the method is
obviously a natural extension. It was called a Lobatto-Galerkin method [Young (1981)], a name
which caused subsequent work to reference it as just another run of the mill finite element
paper. The other names hp Spectral Element method and G-NI or SEM-NI (Galerkin or
Spectral Element Method with Numerical Integration) also fail to highlight the method’s most
important feature, viz. that quadratures are not required. Whatever name is used it should
include the word “collocation”, which is synonymous with “no integration required”. The method
is equivalent to collocation. Even the element interface condition corresponds to setting
residuals to zero, see Eq. (64). By all rights it should be called the C0 Orthogonal Collocation
Finite Element Method, while the alternate collocation at Gauss points should be called the C1
Orthogonal Collocation Finite Element Method. These names include the important word
“collocation” and honor the original one given by Villadsen and Stewart. These names will be
used for the remainder of this article.

6.4 Convergence Rate


Several theoretical convergence results have been proven for the orthogonal collocation Finite
Element methods. Dupont (1976) has summarized the results for the C0 Galerkin method and
the H1 Galerkin method which is equivalent to moments. The convergence and
superconvergence rates for the orthogonal collocation methods are the same as for the finite
element methods they approximate. The rates for C1 orthogonal collocation are the same as
for the H1 Galerkin method or moments, while those for the C0 orthogonal collocation method
are the same as for a C0 Galerkin method with Lagrange trial functions. The overall
convergence rate, e.g. L2 norm, is O(Δxn+2) which is optimal since n+1 is the polynomial
degree. However, they also exhibit some superconvergence properties. C1 orthogonal
collocation converges with O(Δx2n+2) for both the solution and its first derivative, i.e. fluxes, at
element boundary nodes [DeBoor and Swartz (1973), Douglas and Dupoint (1973)]. The C0
method exhibits the same overall convergence and superconvergence rates with less stringent
conditions on smoothness [Diaz (1975), Dunn and Wheeler (1976), Wheeler (1977)]. It should
be no surprise that to achieve this rate the flux must be calculated by the method described in

[45]
section 3.4, i.e. using Eq. (60) evaluated at the element boundary. It has also been shown that
for the C0 method the solution at the Lobatto points within the element is O(Δxn+3) [Nakao
(1984)]. These estimates are based on certain assumptions which will not apply in all cases,
but they indicate the potential accuracy of these methods. The convergence and
superconvergence results have been demonstrated in some of the references above and by
Carey, et al. (1981). Due to the equivalence demonstrated, these convergence rates apply to
all the C0 collocation methods discussed in section 6.3, regardless of the name given.

These convergence rates are in terms of element size, Δx. In hp finite element terminology, this
refinement is call h refinement. The other alternative is to increase the degree of the
polynomial trial function, i.e. p refinement. Maday and Patera (1989) analyzed this alternative
and confirmed exponential convergence is achieved with the C0 orthogonal collocation or hp
spectral element method. Actually, an exponential convergence rate is implied by the O(Δxn+2),
since for increasing n, the rate is not proportional to any fixed power of Δx. Apparently, the
superconvergence properties of this method have been overlooked in the spectral literature.

6.5 Efficiency
The improved efficiency of collocation at Gauss points relative to a conventional Galerkin
method is well documented [Young (1977), Houstis, et al. (1978), Finlayson (1979), Frind and
Pinder (1979)]. A Hermite cubic or higher order grid like Fig. 19 would likely be more efficient
than explicitly imposing continuity with Eq. (58), especially in multiple dimensions. These trial
functions reduce the dimension of the problem substantially, but there is some loss of
efficiency since the nodes and collocation points no longer coincide. Placing single nodes at
Gauss points when possible, as in Fig. 19, should realize some economy since at least some
of the nodes and collocation points are coincident. Collocation at Gauss points saves some
computations relative to integrated moments, but the savings are not nearly as great as with
global methods.

These methods are easily extended to multiple dimensions and trial functions of any order. A
two-dimensional extension of C1 orthogonal collocation was described by Prenter and Russell
(1976). The method has also been extended to distorted
grids and has found many applications in water resource
modeling [Pinder, et al. (1978), Frind and Pinder (1979),
Lapidus and Pinder (1999)]. However, the element
distortion is more restrictive than with a conventional C0
isoparametric finite element method.

When the same trial functions are used, both finite


element moments and collocation at Gauss points provide
substantial savings over the Galerkin method, especially
Fig. 21 Domains of influence,
in multiple dimensions. Fig. 21 shows the domains of - - - - Moments or Collocation
influence for Hermite cubic trial functions in two ______
Galerkin
from Young (1977)

[46]
dimensions. Since the weight functions are local to each element, the approximating matrix
has only 16 nonzero entries per row for collocation or moments rather than 36 for the Galerkin
method. One study found collocation to be just as accurate as the Galerkin method [Frind and
Pinder (1979)]. Another comparison for a convective-diffusion problem found that a Galerkin
method is more accurate than collocation but only slightly more accurate than moments
[Young (1977)]. When the computational complexity was considered, moments was the most
attractive of these methods. The moments method warrants more consideration than it has
received in the past.

The C0 orthogonal collocation method also extends easily to multiple dimensions using tensor
product Lagrange trial functions. It has the following attractive computational features:
1. no quadratures to perform
2. nodes and quadrature points coincide
3. lumped mass or capacitance matrix
4. symmetric matrices for self-adjoint operators
5. can be used with conventional isoparametric elements
6. computational molecules with star structure for rectangular grids

Many of these features are lost in the hybrid formulation.


Since the nodes and quadrature points coincide, no
interpolations are required to calculate coefficients. The
lumped mass matrix permits the use of explicit time
stepping methods. More efficient solution techniques can
be used for symmetric matrix problems.

Fig. 22 shows the computational molecules for grids of


rectangular elements when no cross terms appear in the
differential equation. The linear case is equivalent to the 5-
point finite difference method. With quadratic, cubic and
quartic elements the average number of points in the
difference equations are 7, 9 and 11, respectively. The
cubic method is a true fourth order method with the same
number of neighboring nodes as a 9-point finite difference
or linear finite element method.

Several studies have found this procedure to be very


efficient relative to conventional Galerkin finite element
methods with full integration [Gray (1977), Young (1977),
Melenk, et al. (2001), Schillinger, et al. (2015)]. The
method has found many applications in computational
fluid dynamics [Karniadakis and Sherwood (2013)] and Fig. 22 Difference stencils in 2D
- - - - C0 Orthogonal Collocation
seismic wave propagation [Komatitsch and Villote (1998)]. ______
C0 Galerkin
from Young (1977)

[47]
In one study, the computational characteristics of the C0 orthogonal collocation method were
found to be effective for solving convective-diffusion problems in two spatial dimensions with a
combination of explicit and implicit time stepping [Young (1978, 1981)]. Explicit time stepping
was efficient due to lumping and to the small computational molecules. The implicit equations
were solved with Line Successive Overrelaxation (LSOR) with an optimum relaxation
parameter calculation [Varga (2009)]. Compared to the linear case, equivalent to 5-point finite
differences, the iterations required per node are only 10, 20 and 25 percent greater for
quadratic, cubic and quartic elements, respectively. Due to the difference stencils shown in
Fig. 22, the computational effort per node with cubics was roughly equivalent to that for a 9-
point finite difference or linear finite element method. However, the accuracy per node is
substantially greater and produces a method which is more efficient overall. The method was
also compared to a cubic Hermite Galerkin method and it gave similar accuracy with far less
calculation. This code was developed as a true hp finite element method and refinements were
made both by increasing the number of elements and by increasing the degree of the trial
functions. However, quartic and higher order elements were no more efficient than cubics.

The C0 orthogonal collocation finite element method can be used with conventional
isoparametric elements [Gray (1977), Gray and Van Genuchten (1978) ), Karniadakis and
Sherwood (2013)] in two and three dimensions. It has also been extended to triangular and
tetrahedral elements providing even greater flexibility [Dubiner (1991), Sherwin and
Karniadakis (1995), Cohen, et al. (2001), Mulder (2001)]. All of these developments are based
on Villadsen and Stewart’s idea of placing collocation points at quadrature base points.

Summary
Orthogonal Collocation, Pseudospectral and Differential Quadrature methods are the same, so
this review and its conclusions applies to all. Lanczos’ 80 year old idea was to collocate at the
roots of Chebyshev orthogonal polynomials, avoiding the Runge phenomenon. His idea was
refined by Villadsen and Stewart’s 51-year-old idea to collocate at Lobatto or Gaussian
quadrature base points, also the roots of orthogonal polynomials. The idea of collocating at
quadrature base points has proven useful for extending the method to finite element
approximations.

This retrospective:
1. Reviews development of the method in the context of numerical integration of the
Methods of Weighted Residuals;
2. Shows the Galerkin method is accurately approximated by collocation at Lobatto points;
3. Shows equivalence of: moments and tau methods, accurately approximated by
collocation at Gauss points;
4. Illustrates a simple modified formulation which gives more efficient symmetric matrix
problems for self-adjoint operators with Gauss, Lobatto or Radau points;

[48]
5. Provides compelling support for the correct, weak or natural formulation of flux boundary
conditions with Lobatto and Chebyshev points;
6. Shows how the idea of collocating at quadrature points extends naturally to:
a. C1 Orthogonal Collocation Finite Element method – collocating at Gauss points
b. C0 Orthogonal Collocation Finite Element method – collocating at Lobatto points
7. Demonstrates equivalence of the C0 Orthogonal Collocation Finite Element method with:
a. Hybrid-Collocation-Galerkin method [Diaz (1975), Dunn and Wheeler (1976), Wheeler
(1977)]
b. Lobatto-Galerkin method [Gray (1977), Young (1977, 1981), Hennart (1982)]
c. hp Spectral Element method [Maday and Patera (1989), Karniadakis and Sherwood
(2013)]

Notation
a – trial function coefficients, solution coefficients Eq. (1)
A – first derivative differentiation matrix defined by Eq. (20)
B – 2nd derivative differentiation matrix defined by Eq. (20)
Bi – Biot number, dimensionless external mass transfer coefficient
c – coefficients of mixing cup average composition, Eq. (51)
C – stiffness matrix defined by Eqs. (29) and (34)
d – coefficients of flux, Eq. (52)
D – mass matrix defined by Eq. (28)
f – generic function
g – generic function
ℓ - Lagrange interpolating polynomial defined by Eq. (2)
ℓ* - Lagrange polynomial through interior points only, Eq. (25)
m – number of quadrature points
n – number of interior points
ne – number of elements
P – Jacobi polynomial
r – reaction rate, source function
r̂ - normalized source function
R – differential equation residual
Sh – Sherwood number, dimensionless internal mass transfer coefficient, Eq. (43)
v – eigenfunctions
w – weight or test function in MWR
W – quadrature weight
x – coordinate
y – dimensionless composition
y* - exact solution
y̅ - mixing cup average composition, Eq. (42)
z – axial coordinate

[49]
Greek Letters
α– parameter in Jacobi polynomial weight function, Eq. (4)
β – parameter in Jacobi polynomial weight function, Eq. (4)
δ – Dirac delta or Kronecker delta
ϵp – definition of Lp error, Eq. (23)
ϵη – error in η, Eq. (24)
η – effectiveness factor, normalized flux, Eqs. (13) or (14)
λ – eigenvalue
ξ – local element coordinate
τ – coefficient in residual, Eq. (9)
φ – Thiele modulus
ψ – generic trial or basis function

References
Ames, W. F., Nonlinear partial differential equations in engineering, Academic Press (1965)
Anderson, E., Bai, Z., Bischof, C., Blackford, S., Demmel, J., Dongarra, J., Du Croz, J.,
Greenbaum, A., Hammarling, S., McKenney, A., Sorensen, D., LAPACK Users' Guide (Third
ed.) SIAM, Philadelphia, PA (1999).
Bellman, R. and Casti, J.: “Differential Quadrature and Long-Term Integration,” J. Math. Anal.
Appl., 134, 235-238 (1971).
Bellman, R., Kashef B.G., and Casti, J.: “Differential Quadrature: A Technique for the Rapid
Solution of Nonlinear Partial Differential Equations,” J. Comp. Physics 10, 40-52 (1972).
Bellman, R., Methods of Nonlinear Analysis, 2 Academic Press, New York (1973).
Bellomo, N.: “Nonlinear Models and Problems in Applied Sciences from Differential Quadrature
to Generalized Collocation Methods,” Math. Comput. Modelling, 26, 4, 13-34 (1997).
Bernardi, C. and Maday, Y.: “Spectral Methods”, in Handbook of Numerical Analysis Vol. 5, in
Techniques of Scientific Computing (Part 2), Ciarlet & Lions (ed.) (1997).
Bert, C.W. and Malik, M.: “Differential Quadrature Method in Computational Mechanics: A
Review,” Appl. Mech. Review, 49, 1 (1996).
Bird, R.B., Stewart, W.E. and Lightfoot, E.N., Transport Phenomena, John Wiley and Sons,
(1960).
Boyd, J.P., Chebyshev and Fourier Spectral Methods, 2nd ed., Dover Publications (2000).
Bubnov, I.G.: “Report on the works of Professor Timoshenko which were awarded the
Zhuranskyi Prize,” Symposium of the Institute of Communication Engineers, No. 81, All Union
Special Planning Office (1913).
Canuto, C. and Quarteroni, A.: “Spectral and Pseudo-spectral Methods for Parabolic Problems
with Nonperiodic Problems Boundary Conditions,” Calcolo, 18, 3 197-217 (1981).

[50]
Canuto, C., Boundary Conditions in Chebyshev and Legendre Applications, NASA Contract
Report 172469 (1986)
Canuto, C., Hussaini, M., Quarteroni, A. and Zang, T., Spectral Methods in Fluid Dynamics,
Springer, Berlin (1988)
Canuto, C., Hussaini, M., Quarteroni, A. and Zang, T., Spectral Methods: Fundamentals in
Single Domains, Springer, Berlin (2006)
Canuto, C., Hussaini, M., Quarteroni, A. and Zang, T., Spectral Methods Evolution to Complex
Geometries and Applications to Fluid Dynamics, Springer, Berlin (2007)
Carey, G. and Finlayson, B.A.: “Orthogonal Collocation on Finite Elements”, Chem. Engr. Sci.
30, 587-596 (1975).
Carey, G., Humphrey, D., and Wheeler, M. F.: “Galerkin and Collocation-Galerkin Methods
with Superconvergence and Optimal Fluxes,” International Journal for Numerical Methods in
Engineering, 17, 6, 939-950 (1981).
Ciarlet, P.G. and Raviat, P.A.: “The Combined Effect of Curved Boundaries and Numerical
Integration in Isoparametric Finite Element Methods,” in The Mathematical Foundations of the
Finite Element Method with Applications to Partial Differential Equations, A.K. Aziz, ed.,
Academic Press, New York, pp. 409–474 (1972).
Ciarlet, P.G. The Finite Element Method for Elliptic Problems, North-Holland, Amsterdam,
(1978).
Clenshaw, C.W. and Curtis, A.R.: “A Method for Numerical Integration and an Automatic
Computer,” Numerische Mathematik, 2, 197 (1960).
Clenshaw, C. W. and Norton, H. J.: “The Solution of Nonlinear Ordinary Differential Equations
in Chebyshev Series," Comp. J., 6, 88-92 (1963).
Cohen, G., Joly, P. , Roberts, J. E. and Tordjman, N.: “Higher Order Triangular Finite Elements
With Mass Lumping for the Wave Equation,” SIAM J. Numer. Anal., 38, 6, 2047-2078 (2001).
Collatz, L., The Numerical Treatment of Differential Equations, Springer, 1960
Courant, R., remarks on “weighted averages of the residual”, in discussion following - Biezeno,
C.B. “Graphical and Numerical Methods for Solving Stress Problems,” Proc. 1st Intl. Congress
Appl. Mech., Delft (1924), p. 17.
Crandall, S. H., Engineering Analysis, McGraw-Hill, (1956).
Davis, M.E.: Numerical Methods and Modeling for Chemical Engineers, John Wiley, New York
(1984)
DeBoor, C. and Swartz, B.: “Collocation at Gaussian Points”, SIAM J. Numer. Anal., 10 582–
606 (1973).
Diaz, J., A Hybrid Collocation-Galerkin Method for Two-point Boundary Value Problems Using
Continuous Piecewise Polynomial Spaces, Ph.D. Thesis, Rice Univ. (1975)

[51]
Douglas, J., Jr., Dupont, T.: “A Finite Element Collocation Method for Quasilinear Parabolic
Equations”, Math. Comp., 27 17-28 (1973).
Douglas, J., Jr., Dupont, T. and Wheeler, M. F.: “H1-Galerkin Methods for the Laplace and
Heat Equations”, Math. Aspects of Finite Elements in Partial Differential Equations, Carl De
Boor, Ed., Academic Press, 383 (1974).
Dubiner, M.: “Spectral Methods on Triangles and Other Domains,” J. of Sci. Comp., 6, 4, 345–
390 (1991).
Dunn, R., and Wheeler, M. F.: “Some Collocation-Galerkin Methods for Two-Point Boundary
Value Problems,” SIAM J. Numer. Anal., 13, 5, 720-733 (1976).
Dupont, T.: “A Unified Theory of Superconvergence for Galerkin Methods for Two-Point
Boundary Problems”, SIAM J. Numer. Anal., 13, 3, 362–368 (1976).
Dyksen, W.R., Houstis, E. N., Lynch, R. E. and Rice, J. R.: “The Performance of the
Collocation and Galerkin Methods with Hermite Bi-cubics,” SIAM J. Num. Anal., 21 4, 695–715
(1984).
El Daou, M., Ortiz, E.L., Samara, H.: “A Unified Approach to the Tau Method and Chebyshev
Series Expansion Techniques,” Computers Math. Applic., 25, 3, 73-82 (1993).
El Daou, M., Ortiz, E.L.: "The Tau Method as an Analytic Tool in the Discussion of Equivalence
Results Across Numerical Methods," Computing, 60, 365–376 (1998).
Elnashaie, S.S.E. and Cresswell, D.L.: “On Dynamic Modelling of Porous Catalyst Particles,”
Chem. Engr. Sci. 28, 1387 (1973).
Faddeeva, V.N.: Computational Methods of Linear Algebra, Dover Publications, New York
(1959).
Ferguson, N.B., Orthogonal Collocation as a Method of Analysis in Chemical Reaction
Engineering, Ph.D. thesis, University of Washington, Seattle, WA (1971).
Ferguson, N.B. and Finlayson, B.A.: “Error Bounds for Approximate Solutions to Nonlinear
Ordinary Differential Equations,” Amer. Inst. Chem. Engr. J., 18 (5), 1053-1059 (1972).
Finlayson, B.A. and Scriven, L.E.: “The Method of Weighted Residuals – A Review,” Appl.
Mech. Review. 19, 9, 735-748 (1966).
Finlayson, B.A., The Method of Weighted Residuals and Variational Principles, Academic
Press, New York, NY (1972).
Finlayson, B.A.: “Orthogonal Collocation in Chemical Reaction Engineering,” Cat. Rev. Sci-
Eng. 10, 69-138 (1974).
Finlayson, B.A.: “Orthogonal Collocation on Finite Elements – Progress and Potential,”
Advances in Computer Methods for Partial Differential Equations-III, R. Vichnevetsky (ed.),
IMACS(AICA), Rutgers U., New Brunswick, N.J., 107-113 (1979)

[52]
Finlayson, B.A., Nonlinear Analysis in Chemical Engineering, Ravenna Park Publ., Seattle
(2003).
Fix, G.J.: “Effect of Quadrature Errors in the FInite Element Approximation of Steady State,
Eigenvalue and Parabolic Problems,” in The Mathematical Foundations of the Finite Element
Method with Applications to Partial Differential Equations, A.K. Aziz (ed.), Academic Press, pp.
525–556 (1972).
Franklin, J., Matrix Theory, Prentice-Hall, Englewood Cliffs, NJ (1968).
Frazer, R. A, Jones, W. P., and Skan, S. W.: "Approximation to Functions and to the Solutions
of Differential Equations," Great Britain Aero. Res. Council, London, Report and Memo No.
1799, (1937).
Fried, I.: “Numerical Integration in the Finite Element Method,” Computers & Structures, 4, 5,
921–932 (1974).
Fried, I. and Malkus, D.S.: “Finite Element Mass Matrix Lumping by Numerical Integration with
No Convergence Rate Loss,” Intl. J. Solids and Structures, 11, 4, 461–466 (1975).
Frind, E.O. and Pinder, G.F.: “A Collocation Finite Element Method for Potential Problems in
Irregular Domains,” Int. J. Num. Methods Engr., 14 681-701, (1979)
Fujita, H.: “Diffusion with a Sharp Moving Boundary,” J. Chem. Phys. 21, 700-705 (1953).
Funaro, D., Polynomial Approximation of Differential Equations, Lecture notes in Physics, m8,
Springer, Heidelberg (1992).
Galerkin, B.G.: “Series Solution of Some Problems Related to Elastic Equilibrium of Rods and
Plates,” Vestn. Inzhenerov Tech., 19, 897-908 (1915).
Golub, G.H. and Welsch, J.H.: “Calculation of Gaussian Quadrature Rules”, Mathematics of
Computation, 23, 221-230 (1969).
Gottlieb, D. and Lustman, L.: “The Dufort-Frankel Method for Parabolic Initial Boundary Value
Problems”, Computers and Fluids, 11, 2, 107-120 (1983).
Gottlieb, D. and Orzag, S.A., Numerical Analysis of Spectral Methods: Theory and
Applications, SIAM, Philadelphia, PA (1977).
Gray, W.G.: “An Efficient Finite Element Scheme for Two-Dimensional Surface Water
Computations”, Finite Elements in Water Resources, W.G. Gray, G.F. Pinder and C.A. Brebbia
(ed.), Pentech Press, London (1977).
Gray, W.G. and Van Genuchten, M.T.: “Economical Alternatives to Guassian Quadrature Over
Isoparameteric Quadralaterals,” Intl. J. Num. Meth. Eng. 12 1478-1484 (1978).
Hairer, E., Nörsett, S.P. and Wanner, G., Solving Ordinary Differential Equations I: Nonstiff
Problems, 2nd ed., Springer Verlag, Berlin (1993).
Hairer, E. and Wanner, G., Solving Ordinary Differential Equations II: Stiff and Differential-
Algebraic Problems, 2nd ed., Springer Verlag, Berlin (1996).

[53]
Hennart, J.-P.: “Topics in Finite Element Discretization of Parabolic Evolution Problems,”
Lecture Notes in Math. 909, Springer-Verlag, Berlin, Heidelberg, (1982). [Cohen, et al. (2001)
list an earlier reference which could not be located: Hennart, J.-P., Sainz, E., and Villegas, M.:
“On the Efficient Use of the Finite Element Method in Static Neutron Diffusion Calculations,”
Computational Methods Nuclear Engrg., 1, 3.87 (1979)]
Hildebrand, F.B., Introduction to Numerical Analysis, 2nd ed., Dover Publications, New York,
NY (1987).
Houstis, E.N., Lynch, R.E., Rice, J.R. and Papatheodorou, T.S.: “Evaluation of Numerical
Methods for Elliptic Partial Differential Equations,” J. Comp. Phys., 27, 3, 323–350 (1978).
Hughes, T.J.R., The Finite Element Method: Linear Static and Dynamic Finite Element
Analysis, Prentice-Hall (1987).
Johnson, D.: Chebyshev Polynomials in the Spectral tau Method and Applications to
Eigenvalue Problems, NASA Contractor Report No. 198451 (1996).
Kantorovich, L.V.: “On an Approximation Method for the Solution of a Partial Differential
Equation”, Dokl. Akad. Nauk SSSR, 2, 532-536 (1934)
Kantorovich, L.V. and Krylov, V.I, Approximate Methods in Higher Analysis, Interscience
(1958).
Karniadakis, G. and Sherwin, S.: Spectral/hp Element Methods for Computational Fluid
Dynamics, Oxford Univ. Press (2013).
Karpilovskaya, E.B.: “Convergence of the Collocation Method for Certain Boundary Value
Problems of Mathematical Physics,” Sibirsk. Matem. Zh., 4, 3, 632-640 (1963).
Komatitsch, D. and Villote, J.-P.: “The Spectral Element Method: An Efficient Tool to Simulate
the Seismic Response of 2D and 3D Geologic Structures,” Bull. Seismological Soc. America,
88, 2, 368-392 (1998)
Kravchuk, M.P.: “On Krylov’s Method in the Theory of Approximate Integration of Differential
Equations,” (in Ukrainian),Tr. Fiz. – Mat. Viddilu VUAN, 5, 2, 12-33 (1926) [reference from
Lucka and Lucka (1992)]
Kravchuk, M.P.: “Application of the Method of Moments to the Solution of Linear Differential
and Integral Equations,” (in Ukrainian), Kiev. Soobshch. Akad. Nauk UkSSR, 1, 168 (1932);
[referenced in Kantorovich and Krylov (1958), Mikhlin (1964), Shuleshko (1959)].
Křížek, M. and Neittaanmäki, P.: “Bibliography of Superconvergence”, in Finite Element
Method, Superconvergence, Post-Processing and a Posteriori Estimates, 196, 315-348 Marcel
Dekker, New York (1998).
Krylov, N.M.: “On a Method of Approximate Integration for Which the Ritz Method as Well as
the Method of Least Squares are Special Cases,” C.R. Acad. Sci. Paris, 182, 676-678 (1926)
[reference from Lucka and Lucka, 1992].

[54]
Krylov, N.M.: Collected Works in 3 volumes (in Ukranian) , Izd. Akad. Nauk. UkrSSR, Kiev, vol.
1-3, (1961) [reference from Lucka and Lucka, 1992].
Krylov, V.I., Approximate Calculation of Integrals, Macmillan, New York, NY (1962), Dover
(2012).
Lanczos, C.: “Trigonometric Interpolation of Empirical and Analytical Functions”, J. Math.
Phys., 17, 123-199 (1938).
Lanczos, C., Applied Analysis, Prentice-Hall, Englewood Cliffs, NJ (1956).
Lanczos, C.: “Legendre Versus Chebyshev Polynomials,” in Topics in Numerical Analysis,
Proceedings of the Royal Irish Academy Conference on Numerical Analysis 1972, J.J.H. Miller,
ed., Academic Press, London, pp. 191–201 (1973).
Lapidus, L. and Pinder, G.F.: Numerical Solution of Partial Differential Equations in Science
and Engineering, John Wiley, New York (1999)
Leyk, Z.: “A Co-Collocation-Like Method for Boundary Value Problems,” Numerische
Mathametik, 49, 39-53 (1986).
Leyk, Z.: “A Co-Collocation-Like Method for Elliptic Equations on Rectangular Regions,” J.
Austral. Math. Soc. Ser. B , 38, 368-387 (1997).
Lucka, T.F. and Lucka, A.Y.: “Development of Direct Methods in Mathematical Physics in the
Works of M.P. Kravchuk,” translated from Ukrain. Math. J., 44, 7, 931-939 (1992).
Maday, Y. and Patera, A. T.: “Spectral Element Methods for the Incompressible Navier-Stokes
Equations” In State-of-the-Art Surveys on Computational Mechanics, A.K. Noor, editor, ASME,
New York (1989).
Melenk, J.M., Gerdes, K. and Schwab, C.: “Fully Discrete hp-Finite Elements: Fast
Quadrature,” Comp Meth. in Appl. Mech. and Eng., 190, 32, 4339–4364 (2001).
Michelsen, M.L. and Villadsen, J.V.: "A Convenient Computational Procedure for Collocation
Constants", Chemical Engineering J, 4, 64-68 (1972).
Michelsen, M.L. and Villadsen, J.V.: “Polynomial Solution of Differential Equations”, in
Foundations of Computer-Aided Chemical Process Design, Proceedings of an International
Conference (pp. 341-368), Mah, R.S.H. and Seider, W.D. ed. (1981).
Mikhlin, S. G., Variational Methods in Mathematical Physics, Pergamon/Macmillan (1964).
Mulder, W.A.: “Higher-Order Mass-Lumped Finite Elements for the Wave Equation,” J.
Comput. Acoustics, 9, 2, 671-680 (2001).
Nakao, M.: “Superconvergence Estimates at Jacobi Points of the Collocation-Galerkin Method
for Two Point Boundary Value Problems”, J. Information Processing, 7 31-34 (1984).
Nielson, K.L.: Methods in Numerical Analysis, MacMillan, NY, pp. 150-4 (1956).
Ortiz, E.L.: “Step by Step tau Method – Part I. Piecewise Polynomial Approximations,” Comput.
and Math. Applic., 1, 381-392 (1975).

[55]
Orzag, S.A.: “Comparison of Pseudo Spectral and Spectral Approximation,” Studies in Applied
Mathematics 51, 3, 253-259 (1972).
Orszag, S.A.: “Spectral Methods for Problems in Complex Geometries,” J. of Comp. Physics,
37 1, 70–92 (1980).
Peyret, R.: Spectral Methods for Incompressible Viscous Flows, Springer (2002).
Picone, M.: “Sul Metodo Delle Minime Potenze Ponderate e Sul Metodo di Ritz per il Calcolo
Approssimato nei Problemi Della Fisica-Matematica,” Rend. Circ. Mat. Palermo, 52, 225-253
(1928).
Pinder, G.F., Frind, E.O. and Celia, M A.: “Groundwater Flow Simulation Using Collocation
Finite Elements,” Proc. Second Int. Conf. Finite Elements in Water Resources, (eds Brebbia et
al.), 1.171-1.185, Pentech Press, London, (1978).
Prenter, P.M. and Russell, R.D.: ”Orthogonal Collocation for Elliptic Partial Differential
Equations”, SIAM J. Numer. Anal., 13, 923 (1976).
Quan, J.R. and Chang, C.T.: “New Insights in Solving Distributed System Equations by the
Quadrature Method – I. Analysis,” Comput. Chem. Engng., 13, 7, 779-788 (1989).
Raviart, P.A.: “The Use of Numerical Integration in Finite Element Methods for Parabolic
Equations,” in Topics in Numerical Analysis, Proceedings of the Royal Irish Academy
Conference on Numerical Analysis 1972, J.J.H. Miller, ed., Academic Press, London, pp. 233–
264 (1973).
Ritz, W.: “Über eine neue Methode zur Lösung gewisser Variationsprobleme der
mathematischen”, Physik, J. reine angew. Mathematik 135 1-61 (1908), also collected works,
Paris (1911).
Runge, M.: “Über Empirische Functionen und die Interpolation Zwischen Aequidistanten
Ordinaten,” Zeitschrift für Math. und Physik tome XLVI (1901).
Schillinger, D., Evans, J.A., Frischmann, F., Hiemstra, R.R., Hsu, M.-C., and Hughes, T.J.R.:
“A Collocated C0 Finite Element Method: Reduced Quadrature Perspective, Cost Comparison
with Standard Finite Elements, and Explicit Structural Dynamics,” Intl. J. Num. Meth. Engrg.,
102, no. 3-4, 576-631 (2015).
Shen,J., Tang,T. and Wang,L.-L., Spectral Methods: Algorithms, Analysis and Applications,
Springer-Verlag, Series in Computation Mathematics, vol. 41, Berlin (2011).
Sherwin, S.J. and Karniadakis, G.E.: “A New Triangular and Tetrahedral Basis for High-Order
(hp) Finite Element Methods,” Intl. J. Num. Meth. in Engr., 38, 22, 3775–3802 (1995).
Shu, C., Differential Quadrature and Its Application in Engineering, Springer, London, 2000
Shuleshko, P.: “A New Method of Solving Boundary-Value Problems of Mathematical Physics,
Generalizations of Previous Methods of Solving Boundary-Value Problems of Mathematical
Physics,” Austral. J. Appl. Sci. 10, 1-7, 8-16 (1959).
Slater, J. C.: "Electronic Energy Bands in Metal," Phys. Rev., 45, 794-801 (1934).
[56]
Sorensen, J.P., Guertin, E.W., and Stewart, W.E.: “Computational Models for Cylindrical
Catalyst Particles,” Amer. Inst. Chem. Eng. J., 19, 5, 969-975 (1973).
Strang, G. and Fix, G.J.: An Analysis of the Finite Element Method, Prentice-Hall, 1973
Trefethen, L.N., Spectral Methods in Matlab, SIAM, Philadelphia, PA (2000).
Trefethen, L.N. "Is Gauss Quadrature Better than Clenshaw-Curtis?". SIAM Review 50, 67–87
(2008).
Van de Vosse, F.N. and Minev, P.D.: Spectral elements methods: Theory and applications,
EUT Report 96-W-001, Eidenhoven Univ. Tech., The Netherlands (1996) 96-W-001,
Eindhoven University of Technology, 1996.
Varga, R., Matrix Iterative Analysis, Springer (2009).
Villadsen, J.V. and Stewart, W.E.: “Solution of Boundary-Value Problems by Orthogonal
Collocation,” Chem. Engr. Sci. 22, 1483-1501 (1967).
Villadsen J., Selected Approximation Methods for Chemical Engineering Problems, Inst. for
Kemiteknik Numer. Inst. Danmarks Tekniske Hojskole (1970).
Villadsen, J.V. and Michelsen, M.L., Solution of Differential Equation Models by Polynomial
Approximation, Prentice-Hall, Englewood Cliffs, NJ (1978).
Wheeler, M.F.: “A C0-Collocation-Finite Element Method for Two-Point Boundary Value and
One Space Dimension Parabolic Problems,” SIAM J. Numer. Anal., 14, 1, 71-90 (1977).
Wright, K.: “Chebyshev Collocation Methods for Equations," Comp. J., 6, 358-365 (1964).
Yamada, H., Rept. Res. Inst. Fluid Engng., Kyushu Univ. 3, 29 (1947); referenced by H. Fujita
(1953). Also: “A method of approximate integration of the laminar boundary-layer equation,”
Rept. Res. Inst. Fluid Engng., Kyushu Univ. 6, 87-98 (1950).
Young, L.C. and Finlayson, B.A.: “Mathematical Models of the Monolith Catalytic Converter:
Part I. Development of Model and Application of Orthogonal Collocation,” Amer. Inst. Chem.
Engr. J., 22, 2, 331-343 (1976).
Young, L.C.: “A Preliminary Comparison of Finite Element Methods for Reservoir Simulation,”
Advances in Computer Methods for Partial Differential Equations-II, R. Vichnevetsky (ed.),
IMACS(AICA), Rutgers U., New Brunswick, N.J., 307-320 (1977)
Young, L.C.: “A Finite-Element Method for Reservoir Simulation,” Soc. Petr. Engrs. J. 21(1)
115-128, (Feb. 1981), paper SPE 7413 presented Oct. (1978).
Young, L.C., Orthogonal Collocation Revisited, http://tildentechnologies.com/Numerics (2018).
Zienkiewicz, O.C., The Finite Element Method in Engineering Science, McGraw-Hill (1971).

Appendix A – Galerkin Method with Flux Boundary Conditions

[57]
Consider a problem approximated by Lagrange interpolating polynomials as in Eq. (2). The trial
functions are substituted into the differential equation to form the residual as in Eq. (7). Assume
the boundary conditions are:
𝑦 = 0 at 𝑥 = 0
𝑑𝑦
+ 𝛼𝑦 = 0 at 𝑥 = 1
𝑑𝑥
The usual procedure in orthogonal collocation, pseudospectral or differential quadrature
methods is to satisfy both of these conditions exactly, i.e. use boundary collocation [Finlayson
(1972), p. 101; Villadsen and Michelsen (1978), p. 137; Bert and Malik (1996); Bellomo (1997);
Trefethen (2000), p. 137; Boyd (2000), p. 111, Peyret (2002), p. 59]:
𝑦0 = 0
𝑛+1

∑ 𝐴𝑛+1,𝑖 𝑦𝑖 + 𝛼𝑦𝑛+1 = 0
𝑖=0
where A is the first derivative matrix defined in Eq. (20). After elimination of the boundary
values, the approximation is:
𝑛

𝑦(𝑥) = ∑[ℓ𝑖 (𝑥) + 𝑏𝑖 ℓ𝑛+1 (𝑥)] 𝑦𝑖 (A.1)


𝑖=1
where 𝑏𝑖 = −𝐴𝑛+1,𝑖 /(𝐴𝑛+1,𝑛+1 + 𝛼). When the boundary conditions are satisfied exactly, the
Galerkin method weights the residual, R, by the trial functions which obey the boundary
conditions, i.e. the functions in the brackets of Eq. (A.1). If the Galerkin method integrals are
approximated by quadrature using the n + 2 interior and boundary points, the result is:
𝑛+1

∑ 𝑊𝑘 𝑅(𝑥𝑘 , 𝒚)[ℓ𝑖 (𝑥𝑘 ) + 𝑏𝑖 ℓ𝑛+1 (𝑥𝑘 )] = 𝑊𝑖 𝑅(𝑥𝑖 , 𝒚) + 𝑊𝑛+1 𝑏𝑖 𝑅(1, 𝒚)𝑏𝑖 = 0 (A.2)
𝑘=0
Eq. (A.2) is not equivalent to a collocation method because of the second term involving the
residual at the boundary. Collocation sets only the first interior residual term to zero, so the
nonzero second term seriously violates the Galerkin method criteria. The boundary weight,
Wn+1, is approximately n-2 for Lobatto quadrature, so this inconsistency creates an error over
and above that caused by the approximate quadrature. Numerical experiments suggest the
error introduced is significant. Since the residual at the boundary decreases exponentially with
n, the method with boundary collocation converges exponentially, but usually at a much slower
rate, especially for fluxes. Boundary collocation also causes errors in the overall mass or
energy balance (see Section 3.4).

Boundary collocation works well with Gauss points or Radau points with Wn+1 = 0 (or W0 = 0 if
the flux condition is at x = 0). However, Lobatto or Chebyshev points produce greater accuracy
with a natural treatment of the boundary conditions as described in Section 3.3. Many articles
and texts on pseudospectral and differential quadrature methods state that having a nonzero
boundary weight somehow facilitates the approximation of boundary conditions. If one uses
boundary collocation, the exact opposite is true.

[58]
Appendix B – Links to Available Software
Software useful for implementing MWR and collocation can be found at the following internet
sites:
Funaro (1992) FORTRAN 77 (80 programs, 3500 lines):
morespace.unimore.it/danielefunaro/routines/
Don, Somonoff and Costa Fortran 90 PseudoPack high performance software:
www.cfm.brown.edu/people/wsdon/PseudoPack.htm
Weideman and Reddy (2000) Matlab Differentiation Suite:
appliedmaths.sun.ac.za/~weideman/research/differ.html
Trefethen (2000) software from book:
people.maths.ox.ac.uk/trefethen/spectral.html
Shen, Tang and Wang (2011) approximately 50 Matlab functions from book:
www.ntu.edu.sg/home/lilian/book.htm
Gautschi’s (2005) OPQ Suite, orthogonal polynomials and quadrature:
www.cs.purdue.edu/archives/2002/wxg/codes/OPQ.html
Burkardt (2015), quadrature, numerous other codes in Fortran 90, C++ and Matlab:
www.people.sc.fsu.edu/~jburkardt/
Karniadakis and Sherwood (2013), code in C/C++ oriented toward spectral elements
www.nektar.info/2nd_edition
Finlayson (2003) companion to books, Matlab, Excel, Fortran, Phython, and more
www.chemecomp.com
Young (2018), Matlab/Octave, Fortran 90 and C++ :
www.tildentechnologies.com/Numerics

[59]

View publication stats

You might also like